id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0003/astro-ph0003143.html
ar5iv
text
# 1 Introduction ## 1 Introduction In a thin accretion disk, the time available for the accreting gas to radiate away the energy released by the viscous stress is the accretion time, $$t_{\mathrm{acc}}\frac{1}{\alpha \mathrm{\Omega }_\mathrm{K}}(\frac{r}{H})^2,$$ (1) where $`\alpha `$ is the dimensionless viscosity parameter, $`\mathrm{\Omega }_\mathrm{K}`$ the local Keplerian rotation rate, $`r`$ the distance from the central mass, and $`H`$ the disk thickness (see Frank et al. 1992 or Spruit, elsewhere in this volume). For a thin disk, $`H/r1`$, this time is much longer than the thermal time scale $`t_\mathrm{t}1/(\alpha \mathrm{\Omega })`$. There is then enough time for a local balance to exist between viscous dissipation and radiative cooling. For the accretion rates implied in observed systems the disk is then rather cool, and the starting assumption $`H/r1`$ is justified. This argument is somewhat circular, however, since the accretion time is long enough for effective cooling only if the disk is assumed to be thin to begin with. Other forms of accretion disks may exist, even at the same accretion rates, in which the cooling is ineffective compared with that of standard (geometrically thin, optically thick) disks. Since radiatively inefficient disks tend to be thick, $`H/rO(1)`$, they are sometimes called ‘quasi-spherical’. However, this does not mean that a spherically symmetric accretion model would be a reasonable approximation. The crucial difference is that the flow has angular momentum. The inward flow speed is governed by the rate at which angular momentum can be transferred outwards, rather than by gravity and pressure gradient. The accretion time scale, $`t_{\mathrm{acc}}1/(\alpha \mathrm{\Omega })`$ is longer than the accretion time scale in the spherical case (unless the viscosity parameter $`\alpha `$ is as large as $`O(1)`$). The dominant velocity component is azimuthal rather than radial, and the density and optical depth are much larger than in the spherical case. It turns out that there are two kinds of radiatively inefficient disks, the optically thin, and optically thick varieties. A second distinction occurs because accretion flows are different for central objects with a solid surface (neutron stars, white dwarfs, main sequence stars, planets), and those without (i.e. black holes). I start with optically thick flows. ## 2 Radiation supported advective accretion If the energy loss by radiation is small, the gravitational energy release $`W_{\mathrm{grav}}GM/(2r)`$ is converted into enthalpy of the gas and radiation field<sup>1</sup><sup>1</sup>1I assume here that a fraction $`0.5`$ of the gravitational potential energy stays in the flow as orbital kinetic energy. See also section 3. $$\frac{1}{2}\frac{GM}{r}=\frac{1}{\rho }[\frac{\gamma }{\gamma 1}P_\mathrm{g}+4P_\mathrm{r}],$$ (2) where an ideal gas of constant ratio of specific heats $`\gamma `$ has been assumed, and $`P_\mathrm{r}=\frac{1}{3}aT^4`$ is the radiation pressure. In terms of the virial temperature $`T_{\mathrm{vir}}=GM/(r)`$, and assuming $`\gamma =5/3`$, appropriate for a fully ionized gas, this can be written as $$\frac{T}{T_{\mathrm{vir}}}=[5+8\frac{P_\mathrm{r}}{P_\mathrm{g}}]^1.$$ (3) Thus, for radiation pressure dominated accretion, $`P_\mathrm{r}P_\mathrm{g}`$ , the temperature is much less than the virial temperature. The disk thickness is given by $$H[(P_\mathrm{g}+P_\mathrm{r})/\rho ]^{1/2}/\mathrm{\Omega },$$ (4) With (3) this yields $$H/rO(1).$$ (5) In the limit $`P_\mathrm{r}P_\mathrm{g}`$, the flow is therefore geometrically thick. This implies that radiation pressure supplies a non-negligible fraction of the support of the gas against gravity (the remainder being provided by rotation). For $`P_\mathrm{r}P_\mathrm{g}`$, (2) yields $$\frac{GM}{2r}=\frac{4}{3}\frac{aT^4}{\rho }.$$ (6) The radiative energy flux, in the diffusion approximation, is $$F=\frac{4}{3}\frac{\mathrm{d}}{\mathrm{d}\tau }\sigma T^4\frac{4}{3}\frac{\sigma T^4}{\tau },$$ (7) where $`\sigma =ac/4`$ is Stefan-Boltzmann’s radiation constant. Hence $$F=\frac{1}{8}\frac{GM}{rH}\frac{c}{\kappa }=F_\mathrm{E}\frac{r}{8H},$$ (8) where $`F_\mathrm{E}=L_\mathrm{E}/(4\pi r^2)`$ is the local Eddington flux. Since $`H/r1`$, a radiatively inefficient, radiation pressure dominated accretion flow has a luminosity of the order of the Eddington luminosity. The temperature depends on the accretion rate and the viscosity $`\nu `$ assumed. The accretion rate is of the order $`\dot{M}3\pi \nu \mathrm{\Sigma }`$ (see ‘accretion disks’ elsewhere in this volume), where $`\mathrm{\Sigma }=\rho dz`$ is the surface mass density. In units of the Eddington rate, we get $$\dot{m}\dot{M}/\dot{M}_\mathrm{E}\nu \rho \kappa /c,$$ (9) where $`H/r1`$ has been used, and $`\dot{M}_\mathrm{E}`$ is the Eddington accretion rate onto the central object of size $`R`$, $$\dot{M}_\mathrm{E}=\frac{R}{GM}L_\mathrm{E}=4\pi Rc/\kappa .$$ (10) \[Note that the definition of $`\dot{M}_\mathrm{E}`$ differs by factors of order unity between different authors. It depends on the assumed efficiency $`\eta `$ of conversion of gravitational energy $`GM/R`$ into radiation. In (10) it is taken to be unity, for accretion onto black holes a more realistic value is $`\eta =0.1`$, for accretion onto neutron stars $`\eta 0.4`$.\] Assume that the viscosity scales with the gas pressure: $$\nu =\alpha \frac{P_\mathrm{g}}{\rho \mathrm{\Omega }_\mathrm{K}},$$ (11) instead of the total pressure $`P_\mathrm{r}+P_\mathrm{g}`$. This is the form that is likely to hold if the angular momentum transport is due to a small-scale magnetic field (Sakimoto and Coroniti, 1989). Then with (6) we have (up to a factor $`2H/rO(1)`$ $$T^5\frac{(GM)^{3/2}}{r^{5/2}}\frac{\dot{m}c}{\alpha \kappa a},$$ (12) or $$T\mathrm{2\hspace{0.17em}10}^8r_6^{1/5}(r/r_\mathrm{g})^{3/10}\dot{m}^{1/5},$$ (13) where $`r=10^6r_6`$ and $`r_\mathrm{g}=2GM/c^2`$ is the gravitational radius of the accreting object, and the electron scattering opacity of 0.3 cm<sup>2</sup>/g has been assumed. The temperatures expected in radiation supported advection dominated flows are therefore quite low compared with the virial temperature \[If the viscosity is assumed to scale with the total pressure instead of $`P_\mathrm{g}`$, the temperature is even lower\]. The effect of electron-positron pairs can be neglected (Schultz and Price, 1985), since they are present only at temperatures approaching the electron rest mass energy, $`T>\mathrm{\hspace{0.33em}10}^9`$K. In order for the flow to be radiation pressure and advection dominated, the optical depth has to be sufficiently large so the radiation does not leak out. The energy density in the flow, vertically integrated at a distance $`r`$, is of the order $$EaT^4H,$$ (14) and the energy loss rate per cm<sup>2</sup> of disk surface is given by (8). The cooling time is therefore, $$t_\mathrm{c}=E/F=3\tau H/c.$$ (15) This is to be compared with the accretion time, which can be written in terms of the mass in the disk at radius $`r`$, of the order $`2\pi r^2\mathrm{\Sigma }`$, and the accretion rate: $$t_{\mathrm{acc}}=2\pi r^2\mathrm{\Sigma }/\dot{M}.$$ (16) This yields $$t_\mathrm{c}/t_{\mathrm{acc}}\frac{\kappa }{\pi rc}\dot{M}=4\dot{m}\frac{R}{r},$$ (17) (where a factor $`3/2H/rO(1)`$ has been neglected). Since $`r>R`$, this shows that accretion has to be super-Eddington in order to be both radiation- and advection-dominated. This condition can also be expressed in terms of the so-called trapping radius $`r_\mathrm{t}`$ (e.g. Rees 1978). Equating $`t_{\mathrm{acc}}`$ and $`t_\mathrm{c}`$ yields $$r_\mathrm{t}/R4\dot{m}.$$ (18) Inside $`r_\mathrm{t}`$, the flow is advection dominated: the radiation field produced by viscous dissipation stays trapped inside the flow, instead of being radiated from the disk as happens in a standard thin disk. Outside the trapping radius, the radiation field can not be sufficiently strong to maintain a disk with $`H/r1`$, it must be a thin form of disk instead. Such a thin disk can still be radiation-supported (i.e. $`P_\mathrm{r}P_\mathrm{g}`$), but it can not be advection dominated. Flows of this kind are called ‘radiation supported tori’ (or radiation tori, for short) by Rees et al. 1982<sup>2</sup><sup>2</sup>2Some workers interpret the use of word ‘torus’ in the context radiatively inefficient accretion as implying a rotating but non-accreting flow. The physics studied by Rees et al. (1982), however, where this name was introduced, explicitly refers to accreting flows such as described here. They must accrete at a rate above the Eddington value to exist. The converse is not quite true: a flow accreting above Eddington is an advection dominated flow, but it need not necessarily be radiation dominated. Advection dominated optically thick acretion flows exist in which radiation does not play a major role (see section 3.1). That an accretion flow above $`\dot{M}_\mathrm{E}`$ is advection dominated, not a thin disk, follows from the fact that in a thin disk the energy dissipated must be radiated away locally. Since the local radiative flux can not exceed the Eddington energy flux $`F_\mathrm{E}`$, the mass accretion rate in a thin disk can not significantly exceed the Eddington value (10). The gravitational energy, dissipated by viscous stress in differential rotation and advected with the flow, ends up at the central object. If this is a black hole, the photons, particles and their thermal energy are conveniently swallowed at the horizon, and do not react back on the flow. Radiation tori are therefore mostly relevant for accretion onto black holes. They are convectively unstable (Bisnovatyi-Kogan and Blinnikov 1977): the way in which energy is dissipated, in the standard $`\alpha `$-prescription, is such that the entropy ($`T^3/\rho `$) decreases with height in the disk. Recent numerical simulations (see section 5) show the effects of this convection. ### 2.1 Super-Eddington accretion onto black holes As the accretion rate onto a black hole is increased above $`\dot{M}_\mathrm{E}`$, the trapping radius moves out. The total luminosity increases only slowly, and remains of the order of the Eddington luminosity. Such supercritical accretion has been considered by Begelman and Meier (1982, see also Wang and Zhou 1999); they show that the flow has a radially self-similar structure. Abramowicz et al. (1988, 1989) studied accretion onto black holes at rates near $`\dot{M}_\mathrm{E}`$. They used a vertically-integrated approximation for the disk, but included the advection terms. The resulting solutions were called ‘slim disks’. These models show how with increasing accretion rate, a standard thin Shakura-Sunyaev disk turns into a radiation-supported advection flow. The nature of the transition depends on the viscosity prescription used, and can show a non-monotonic dependence of $`\dot{M}`$ on surface density $`\mathrm{\Sigma }`$ (Honma et al. 1991). This suggests the possibility of instability and cyclic behavior of the inner disk near a black hole, at accretion rates near and above $`\dot{M}_\mathrm{E}`$ (for an application to GRS 1915+105 see Nayakshin et al., 1999). ### 2.2 Super-Eddington accretion onto neutron stars In the case of accretion onto a neutron star, the energy trapped in the flow, plus the remaining orbital energy, settles onto its surface. If the accretion rate is below $`\dot{M}_\mathrm{E}`$, the energy can be radiated away by the surface, and steady accretion is possible. A secondary star providing the mass may, under some circumstances, transfer more than $`\dot{M}_\mathrm{E}`$, since it does not know about the neutron star’s Eddington value. The outcome of this case is still somewhat uncertain; it is generally believed on intuitive grounds that the ‘surplus’ (the amount above $`\dot{M}_\mathrm{E}`$) somehow gets expelled from the system. As the transfer rate is increased, the accreting hot gas forms an extended atmosphere around the neutron star, like the envelope of a giant. If it is large enough, the outer parts of this envelope are partially ionized. The opacity in these layers, due to lines of the CNO and heavier elements, is then much higher than the electron scattering opacity. The Eddington luminosity based on the local value of the opacity is then smaller than it is near the neutron star surface. Once an extended atmosphere with a cool surface forms, the accretion luminosity is thus large enough to drive a wind from the envelope (see Kato 1997, where this is discussed in the context of Novae). This scenario is somewhat dubious however, since it assumes that the mass transferred from the secondary continues to reach the neutron star and generate a high luminosity there. This is not at all obvious, since the mass transfering stream may dissipate in the growing neutron star envelope instead. The result would be a giant (more precisely, a Thorne-Zytkow star), with a steadily increasing envelope mass. Such an envelope is likely to be large enough to envelop the entire binary system, which then develops into a common-envelope (CE) system. The envelope mass is expected to be ejected by CE hydrodynamics (Taam 1994, 2000). A more speculative proposal, suggested by the properties of SS 433, is that the ‘surplus mass’ is ejected in the form of jets. The binary parameters of Cyg X-2 are observational evidence for mass ejection in super-Eddington mass transfer phases (King and Ritter 1999, Rappaport and Podsiadlowski 1999, King and Begelman 1999). ## 3 Hydrodynamics The hydrodynamics of ADAFs and radiation tori can be studied by starting, at a very simple level, with a generalization of the thin disk equations. Making the assumption that quantities integrated over the height $`z`$ of the disk give a fair representation (though this is justifiable only for thin disks), and assuming axisymmetry, the problem reduces to a one-dimensional time-dependent one. Further simplifying this by restriction to a steady flow yields the equations $$2\pi r\mathrm{\Sigma }v_r=\dot{M}=\mathrm{cst},$$ (19) $$r\mathrm{\Sigma }v_r_r(\mathrm{\Omega }r^2)=_r(\nu \mathrm{\Sigma }r^3_r\mathrm{\Omega }),$$ (20) $$v_r_rv_r(\mathrm{\Omega }^2\mathrm{\Omega }_\mathrm{K}^2)r=\frac{1}{\rho }_rp,$$ (21) $$\mathrm{\Sigma }v_rT_rS=q^+q^{},$$ (22) where $`S`$ is the specific entropy of the gas, $`\mathrm{\Omega }`$ the local rotation rate, now different from the Keplerian rate $`\mathrm{\Omega }_\mathrm{K}=(GM/r^3)^{1/2}`$, while $$q^+=Q_\mathrm{v}dzq^{}=\mathrm{div}F_\mathrm{r}dz$$ (23) are the height-integrated viscous dissipation rate and radiative loss rate, respectively. In the case of thin disks, equations (19) and (20) are unchanged, but (21) simplifies to $`\mathrm{\Omega }^2=\mathrm{\Omega }_\mathrm{K}^2`$, i.e. the rotation is Keplerian, while (22) simplifies to $`q^+=q^{}`$, expressing local balance between viscous dissipation and cooling. The left hand side of (22) describes the radial advection of heat, and is perhaps the most important deviation from the thin disk equations at this level of approximation (hence the name advection dominated flows). The characteristic properties are seen most cearly when radiative loss is neglected altogether, $`q^{}=0`$. The equations are supplemented with expressions for $`\nu `$ and $`q^+`$: $$\nu =\alpha c_\mathrm{s}^2/\mathrm{\Omega }_\mathrm{K};q^+=(r_r\mathrm{\Omega })^2\nu \mathrm{\Sigma }.$$ (24) If $`\alpha `$ is taken constant, $`q^{}=0`$, and an ideal gas is assumed with constant ratio of specific heats, so that the entropy is given by $$S=c_\mathrm{v}\mathrm{ln}(p/\rho ^\gamma ),$$ (25) then equations (19)-(22) have no explicit length scale in them. This means that a special so-called self-similar solution exists, in which all quantities are powers of $`r`$. Such self-similar solutions have apparently first been described by Gilham (1981), but have since then been re-invented several times (Spruit et al. 1987; Narayan and Yi, 1994). The dependences on $`r`$ are $$\mathrm{\Omega }r^{3/2};\rho r^{3/2},$$ (26) $$Hr;Tr^1.$$ (27) In the limit $`\alpha 1`$, one finds $$v_r=\alpha \mathrm{\Omega }_\mathrm{K}r\left(9\frac{\gamma 1}{5\gamma }\right),$$ (28) $$\mathrm{\Omega }=\mathrm{\Omega }_\mathrm{K}\left(2\frac{53\gamma }{5\gamma }\right)^{1/2},$$ (29) $$c_\mathrm{s}^2=\mathrm{\Omega }_\mathrm{K}^2r^2\frac{\gamma 1}{5\gamma },$$ (30) $$\frac{H}{r}=\left(\frac{\gamma 1}{5\gamma }\right)^{1/2}.$$ (31) The precise from of these expressions depends somewhat on the way in which vertical integrations such as in (23) are done (which are only approximate). The self-similar solution can be compared with numerical solutions of eqs. (19)–(22) with appropriate conditions applied at inner ($`r_\mathrm{i}`$) and outer ($`r_\mathrm{o}`$) boundaries (Nakamura et al. 1996, Narayan et al. 1997). The results show that the self-similar solution is valid in an intermediate regime $`r_\mathrm{i}rr_\mathrm{o}`$. That is, the solutions of (19)–(22) approach the self-similar solution far from the boundaries, as is characteristic of self-similar solutions. The solution exists only if $`1<\gamma 5/3`$, a condition satisfied by all ideal gases. As $`\gamma 1`$, the disk temperature and thickness vanish. This is understandable, since a $`\gamma `$ close to 1 means that the gas has a large number of internal degrees of freedom. The accretion energy is shared between all degrees of freedom, so that for a low $`\gamma `$ less is available for the kinetic energy (temperature) of the particles. Second, the rotation rate vanishes for $`\gamma 5/3`$. Since a fully ionized gas has $`\gamma =5/3`$, it is the most relevant value for optically thin accretion near a black hole or neutron star. Apparently, steady advection dominated accretion can not have angular momentum in this case, and the question arises how an adiabatic flow with angular momentum will behave for $`\gamma =5/3`$. In the literature, this problem has been circumvented by arguing that real flows would have magnetic fields in them, which would change the effective compressibility of the gas. Even if a magnetic field of sufficient strength is present, however, (energy density comparable to the gas pressure) the effective $`\gamma `$ is not automatically lowered. If the field is compressed mainly perpendicular to the field lines, for example, the effective $`\gamma `$ is closer to 2. Also, this does not solve the conceptual problem what would happen to a rotating accretion flow consisting of a more weakly magnetized ionized gas. This conceptual problem has been solved by Ogilvie (1999), who showed how the low rotation for $`\gamma =5/3`$ comes about in a time-dependent manner. He found a similarity solution (depending on distance and time in the combination $`r/t^{2/3}`$) to the time-dependent version of eqs (19)–(22). This solution describes the asymptotic behavior (in time) of a viscously spreading disk, analogous to the viscous spreading of thin disks (Lynden-Bell and Pringle 1972). As in the thin disk case, all the mass accretes asymptotically onto the central mass, while all the angular momentum travels to infinity together with a vanishing amount of mass. For all $`\gamma <5/3`$, the rotation rate at a fixed $`r`$ tends to a finite value as $`t\mathrm{}`$, but for $`\gamma =5/3`$ it tends to zero. The slowly-rotating region expands in size as $`rt^{2/3}`$. It thus seems likely that the typical slow rotation of ADAFs at $`\gamma `$ near 5/3 is a real physical property, and that the angular momentum gets expelled from the inner regions of the flow. ### 3.1 Other optically thick accretion flows The radiation-dominated flows discussed in section 2 are not the only possible optically thick advection dominated flows. From the discussion of the hydrodynamics, it is clear that disk-like (i.e. rotating) accretion is possible whenever the ratio of specific heats is less than 5/3. A radiation supported flow satisfies this requirement with $`\gamma =4/3`$, but it can also happen in the absence of radiation if energy is taken up in the gas by internal degrees of freedom of the particles. Examples are the rotational and vibrational degrees of freedom in molecules, and the energy associated with dissociation and ionization. If the accreting object has a gravitational potential not much exceeding the 13.6 + 2.2 eV per proton for dissociation plus ionization, a gas initially consisting of molecular hydrogen can stay bound at arbitrary accretion rates. This translates into a limit $`M/M_{}R_{}/R<0.01`$. This is satisfied approximately by the giant planets, which are believed to have gone through a phase of rapid adiabatic gas accretion (e.g. Podolak et al. 1993). A more remotely related example is the core-collapse supernova. The accretion energy of the envelope mass falling onto the proto-neutron star is lost mostly through photodisintegration of nuclei, causing the well known problem of explaining how a shock is produced of sufficient energy to unbind the envelope. If the pre-collapse core rotates sufficiently rapidly, the collapse will form an accretion torus (inside the supernova envelope), with properties similar to advection dominated accretion flows (but at extreme densities and accretion rates, by X-ray binary standards). Such objects have been invoked as sources of Gamma-ray bursts (Popham et al. 1999, see also the review by Meszaros, elsewhere in this volume). A final possibility for optically thick accretion is through neutrino losses. If the temperature and density near an accreting neutron star become large enough, additional cooling takes place through neutrinos (as in the cores of giants). This is relevant for the physics of Thorne-Zytkow stars (neutron stars or black holes in massive supergiant envelopes, cf. Bisnovatyi-Kogan and Lamzin 1984, Cannon et al. 1992), and perhaps for the spiral-in of neutron stars into giants (Chevalier 1993, see however Taam 2000). ## 4 Optically thin advection dominated flows (ADAFs) The optically thin case has received most attention in recent years, because of the promise it holds for explaining the (radio to X-ray) spectra of X-ray binaries and the central black holes in galaxies, including our own. For a recent review see Yi (1999). This kind of flow occurs if the gas is optically thin, and radiation processes sufficiently weak. The gas then heats up to near the virial temperature. Near the last stable orbit of a black hole, this is of the order 100 MeV, or $`10^{12}`$K. At such temperatures, a gas in thermal equilibrium would radiate at a fantastic rate, even if it were optically thin, because the interaction between electrons and photons becomes very strong already near the electron rest mass of 0.5MeV. In a remarkable paper, Shapiro Lightman and Eardley (1976) noted that this, however, is not what will happen in an optically thin accreting plasma but that, instead, a two-temperature plasma forms. Suppose that the energy released by viscous dissipation is distributed equally among the carriers of mass, i.e. mostly to the ions and $`1/2000`$ to the electrons. Most of the energy resides in the ions, which radiate very inefficiently (their high mass prevents the rapid accelerations that are needed to produce electromagnetic radiation). This energy is transfered to the electrons by Coulomb interactions. These interactions are slow, however, under the conditions mentioned. They are slow because of the low density (on account of the assumed optical tickness), and because they decrease with increasing temperature. The electric forces that transfer energy from an ion to an electron act only as long as the ion is within the electron’s Debye sphere (e.g. Spitzer, 1965). The interaction time between proton and electron, and thus the momentum transfered, therefore decrease as $`1/v_\mathrm{p}T_\mathrm{p}^{1/2}`$ where $`T_\mathrm{p}`$ is the proton temperature. In this way, an optically thin plasma near a compact object can be in a two-temperature state, with the ions being near the virial temperature, and the electrons, which are doing the radiating, at a much lower temperature around 50–200 keV. The energy transfer from the gravitational field to the ions is fast (by some form of viscous or magnetic dissipation), from the ions to the electrons slow, and finally the energy losses of the electrons fast (by synchrotron radiation in a magnetic field or by inverse Compton scattering of soft photons). Such a flow would be radiatively inefficient since the receivers of the accretion energy, the ions, get swallowed by the hole before getting a chance to transfer their energy to the electrons. The first disk models which take into account the physics of advection and a two-temperature plasma were developed by Ichimaru (1977). It is clear from this description that both the physics of such flows and the radiation spectrum to be expected depend crucially on the details of the ion-electron interaction and radiation processes assumed. This is unlike the case of the optically thick advection dominated flows, where gas and radiation are in approximate thermodynamic equilibrium. This is a source of uncertainty in the application of the optically thin ADAFs to observed systems, since their radiative properties depend on poorly known quantities, for example, the strength of the magnetic field in the flow. The various branches of optically thin and thick accretion flows are summarized in figure 1. Each defines a relation between surface density $`\mathrm{\Sigma }`$ (or optical depth $`\tau =\kappa \mathrm{\Sigma }`$) and accretion rate. Optically thin ADAFs require low densities, either because of low accretion rates or large values of the viscosity parameter. The condition that the cooling time of the ions by energy transfer to the electrons is longer than the accretion time yields a maximum accretion rate (Rees et al. 1982), $$\dot{m}<\alpha ^2.$$ (32) If $`\alpha 0.05`$ as suggested by current simulations of magnetic turbulence, the maximum accretion rate would be a few $`10^3`$. If ADAFs are to be applicable to systems with higher accretion rates, such as Cyg X-1 for example, the viscosity parameter must be larger, on the order of 0.3. ### 4.1 Application: hard spectra in X-ray binaries In the hard state, the X-ray spectrum of black hole and neutron star accreters is characterized by a peak in the energy distribution ($`\nu F_\nu `$ or $`EF(E)`$) at photon energies around 100 keV. This is to be compared with the typical photon energy of $`1`$ keV expected from a standard optically thick thin disk accreting near the Eddington limit. The standard, and by far most likely explanation is that the observed hard photons are softer photons (around 1 keV) that have been up-scattered through inverse Compton scattering on hot electrons. Fits of such Comptonized spectra (e.g. Zdziarski 1998 and references therein) yield an electron scattering optical depth around unity and an electron temperature of 50–100 keV. The scatter in these parameters is rather small between different sources. The reason may lie in part in the physics of Comptonization, but is not fully understood either. Something in the physics of the accretion flow keeps the Comptonization parameters constant as long as it is in the hard state. ADAFs have been applied with some success in interpreting XRB. They can produce reasonable X-ray spectra, and have been used in interpretations of the spectral-state transitions in sources like Cyg X-1 (Esin et al. 1998 and references therein). An alternative to the ADAF model for the hard state in sources like Cyg X-1 and the black hole X-ray transients is the ‘corona’ model. A hot corona (Bisnovatyi-Kogan and Blinnikov 1976), heated perhaps by magnetic fields as in the case of the Sun (Galeev et al. 1979) could be the medium that Comptonizes soft photons radiated from the cool disk underneath. The energy balance in such a model produces a Comptonized spectrum within the observed range (Haardt and Maraschi 1993). This model has received further momentum, especially as a model for AGN, with the discovery of broadened X-ray lines indicative of the presence of a cool disk close to the last stable orbit around a black hole (Lee et al. 1999 and references therein). The very rapid X-ray variability seen in some of these sources is interpreted as due magnetic flaring in the corona, like in the solar corona (e.g. Di Matteo et al. 1999a). ### 4.2 Transition from thin disk to ADAF One of the difficulties in applying ADAFs to specific observed systems is the transition from a standard geometrically thin, optically thick disk, which must be the mode of mass transfer at large distances, to an ADAF at closer range. This is shown by figure 1, which illustrates the situation at some distance close to the central object. The standard disk and the optically thin branches are separated from each other for all values of the viscosity parameter. This separation of the optically thin solutions also holds at larger distances. Thus, there is no plausible continuous path from one to the other, and the transition between the two must be due to additional physics that is not included in diagrams like figure 1. A promising possibility is that the transition takes place through a gradual (as a function of radius) evaporation (Meyer and Meyer-Hofmeister 1994, Meyer-Hofmeister and Meyer 1999). In this scenario, evaporation initially produces a corona above the disk, which transforms into an ADAF further in. ### 4.3 Quiescent galactic nuclei For very low accretion rates, such as inferred for the black hole in the center of our galaxy (identified with the radio source Sgr A\*), the broad band spectral energy distribution of an ADAF is predicted to have two humps (Narayan et al. 1995, Quataert et al. 1999). In the X-ray range, the emission is due to bremsstrahlung. In the radio range, the flow emits synchrotron radiation, provided that the magnetic field in the flow has an energy density order of the gas pressure (‘equipartition’). Synthetic ADAF spectra can be fitted to the observed radio and X-ray emission from Sgr A\*. In other galaxies where a massive central black hole is inferred, and the center is populated by an X-ray emitting gas of known density, ADAFs would also be natural, and might explain why the observed luminosities are so low compared with the accretion rate expected for a hole embedded in a gas of the measured density. In some of these galaxies, however, the peak in the radio-to-mm range predicted by analogy with Sgr A\* is not observed (Di Matteo et al. 1999b). This requires an additional hypothesis, for example that the magnetic field in these cases is much lower, or that the accretion energy is carried away by an outflow. ### 4.4 Transients in quiescence X-ray transients in quiescence (i.e. after an outburst) usually show a very low X-ray luminosity. The mass transfer rate from the secondary in quiescence can be inferred from the optical emission. This shows the characteristic ‘hot spot’, known from other systems to be the location where the mass transfering stream impacts on the edge of an accretion disk (e.g. van Paradijs and McClintock 1995). These observations thus show that a disk is present in quiescence, while the mass transfer rate can be measured from the brightness of the hot spot. If this disk were to extend to the neutron star with constant mass flux, the predicted X-ray luminosity would be much higher than observed. This has traditionally been interpreted as a consequence of the fact that in transient systems, the accretion is not steady. Mass is stored in the outer parts and released by a disk instability (e.g. King 1995, Meyer-Hofmeister and Meyer 1999) producing the X-ray outburst. During quiescence, the accretion rate onto the compact object is much smaller than the mass transfer from the secondary to the disk. ADAFs have been invoked as an alternative explanation. The quiescent accretion rate onto the central object is proposed to be higher than in the disk-instability explanation, the greater energy release being hidden on account of the low radiative efficiency of the ADAF. Some transient systems have neutron star primaries, with a hard surface at which the energy accreted by the ADAF must somehow be radiated away. A neutron star, with or without ADAFs, can not accrete in a radiatively inefficient way. In order to make ADAFs applicable, it has been proposed that the neutron stars in these systems have a modest magnetic dipole moment, such that in quiescence the gas in the accretion disk is prevented, by the ‘propeller effect’ (Illarionov and Sunyaev 1975, Sunyaev and Shakura 1977) from accreting onto the star. ### 4.5 ADAF-disk interaction: Lithium One of the strong predictions of ADAF models, whether for black holes or neutron stars, is that the accreting plasma in the inner regions has an ion temperature of 10–100 MeV. Nearby is a cool and dense accretion disk feeding this plasma. If only a small fraction of the hot ion plasma gets in contact with the disk, the intense irradiation by ions will produce nuclear reactions (Aharonian and Sunyaev 1984, Martín et al. 1992). The main effects would be spallation of CNO elements into Li and Be, and the release of neutrons by various reactions. In this context, it is intriguing that the secondaries of both neutron star and black hole accreters have high overabundances of Li compared with other stars of their spectral types (Martín et al. 1992, 1994a). If a fraction of the disk material is carried to the secondary by a disk wind, the observed Li abundances may be accounted for (Martín et al. 1994b). ### 4.6 ADAF-disk interaction: hard X-spectra The interaction of a hot ion plasma with the cool disk produces a surface layer heated by the incident ions through Coulomb interactions with the electron gas in the disk. Its thickness and temperature turn out to be largely self-regulating: the energy balance is as in the Haardt and Maraschi corona models (in which the interaction is by photons only), while the optical thickness self-regulates through the dependence of the ion penetration depth on the electron temperature. This model (Spruit 1997) produces hard comptonized X-ray spectra whose shape is largely independent of both the energy flux and distance from the central object. ## 5 Outflows? The energy density in an advection dominated accretion flow is of the same order as the gravitational binding energy density $`GM/r`$, since a significant fraction of that energy went into internal energy of the gas by viscous dissipation, and little of it got lost by radiation. The gas is therefore only marginally bound in the gravitational potential. This suggests that perhaps a part of the accreting gas can escape, producing an outflow or wind. In the case of the ion supported optically thin ADAFs, this wind would be thermally driven by the temperature of the ions, like an ‘evaporation’ from the accretion torus. In the case of the radiation supported tori, which exist only at a luminosity near the Eddington value, but with much lower temperatures than the ion tori, winds driven by radiation pressure could exist. The possibility of outflows is enhanced by the viscous energy transport through the disk. In the case of thin accretion disks (not quite appropriate in the present case, but sufficient to demonstrate the effect), the local rate of gravitational energy release (erg cm<sup>-2</sup>s<sup>-1</sup>) is $`W=\mathrm{\Sigma }v_r_r(GM/r)`$. The local viscous dissipation rate is $`(9/4)\nu \mathrm{\Sigma }\mathrm{\Omega }^2`$. They are related by $$Q_\mathrm{v}=3[1(\frac{r_\mathrm{i}}{r})^{1/2}]W,$$ (33) where $`r_\mathrm{i}`$ is the inner edge of the disk (see ‘accretion disks’ elsewhere in this volume). The viscous dissipation rate is less than the gravitational energy release for $`r<(4/9)r_\mathrm{i}`$, and larger outside this radius. Part of the gravitational energy released in the inner disk is transported outward by the viscous stresses, so that the energy deposited in the gas is up to three times larger than expected from a local energy balance argument. The temperatures in an ADAF would be correspondingly larger. Begelman and Blandford (1999) have appealed to this effect to argue that in an ADAF most of the accreting mass of a disk might be expelled through a wind, the energy needed for this being supplied by the viscous energy transport associated with the small amount of mass that actually accretes. These suggestions are in principle testable, since the arguments are about two-dimensional time dependent flows (axisymmetric), which can be studied fairly well by numerical simulation. Igumenshchev et al. (1996), and Igumenshchev and Abramowicz (1999) present results of such simulations, but unfortunately these give a somewhat ambiguous answer to the question. For large viscosity ($`\alpha 0.3`$) no outflow is seen, but for small viscosity time dependent flows are seen with outflows in some regions. Some of these flows may be a form of convection and unrelated to systematic outflows.
warning/0003/cond-mat0003056.html
ar5iv
text
# Charge-orbital ordering and phase-separation in the two-orbital model for manganites: Roles of Jahn-Teller phononic and Coulombic interactions ## I Introduction The recent discovery of the colossal magnetoresistance (CMR) phenomenon in manganese oxides has triggered a huge experimental and theoretical effort to understand its origin. The large changes in resistivity observed in experiments usually involve both a low-temperature or high magnetic field ferromagnetic (FM) metallic phase, and a high-temperature or low magnetic field insulating phase. For this reason, to address the CMR effect, it appears unavoidable to have a proper understanding of the phases competing with the metallic FM regime, which as a first approximation can be rationalized based on the standard double exchange (DE) mechanism. In fact, the phase diagram of manganites has revealed a very complex structure, clearly showing that metallic ferromagnetism is just one of the several spin, orbital, and charge arrangements that are possible in the manganese oxides. The theoretical study of manganites started decades ago when the DE ideas to explain the FM-phase were proposed. Unfortunately, the model used in those early calculations involved only one orbital, and the many-body techniques used in its analysis were mostly restricted to crude mean-field approximations (MFA). Quite recently, the first computational studies of the one-orbital model for manganites have been presented by Yunoki et al. While the expected FM-phase emerged clearly from such analysis, several novel features were identified, notably phase-separation (PS) tendencies close to $`n`$=1 (where $`n`$ is the $`e_\mathrm{g}`$ electron number density per site) between an antiferromagnetic (AFM) insulating phase and a metallic FM state. Several calculations have confirmed these tendencies toward mixed-phase characteristics, using a variety of techniques. Moreover, the computational work has been extended to include non-cooperative Jahn-Teller (JT) phonons and PS tendencies involving spin-FM phases that differ in their orbital arrangement, i.e., staggered vs. uniform, were again identified using one-dimensional (1D) and two-dimensional (2D) clusters. This illustrates the relevance of the orbital degree of freedom in the analysis of manganites, and confirms the importance of mixed-phase characteristics in this context. The theoretical, mostly computational, calculations have been recently summarized by Moreo et al., where it has been argued that a state with clusters of one phase embedded into another should have a large resistivity and a large compressibility. Other predictions in mixed-phase regimes involve the presence of a robust pseudogap in the density of states, similar to the results obtained using angle-resolved photoemission experiments applied to bilayer manganites. A large number of experiments have reported the presence of inhomogeneities in real manganites, results compatible with those obtained using computational studies. More recently, the influence of disorder on metal-insulator transitions of manganite models that would be of first-order character without disorder has been proposed as the origin of the large cluster coexistence reported in experiments for several compounds. It is clear that intrinsic tendencies toward inhomogeneous states is at the heart of manganite physics, and simple scenarios involving small polarons are not sufficient to understand the mixed-phase tendencies observed experimentally. Although the presence of mixed-phase tendencies competing with the FM-phase is by now well-established and a considerable progress has been achieved in the theoretical study of models for manganites, many issues still remain to be investigated in this context. One of them is related with the notorious charge ordering phenomenon, characteristic of the narrow-band manganites. Especially at $`n`$=0.5, the so-called $`CE`$-type AFM phase has been widely observed in materials such as La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> and Nd<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub>. In spite of the considerable importance of this phase in view of its relevance for the CMR effect reported at high hole density, it is only recently that the $`CE`$-phase has been theoretically understood using the analytic MFA in the JT-phononic model and numerical computational techniques. The importance of the zigzag chains of the $`CE`$-state, along which the spins are parallel, has been remarked in these recent efforts. In fact, it has been shown that even without electron-phonon or Coulomb interactions the $`CE`$-state structure is nevertheless stable, and in this limit it arises from a simple band-insulator picture. Previous work, such as the pioneering results of Goodenough, were based on the $`assumption`$ of a checkerboard charge state upon which the spin and orbital arrangements were properly calculated, but they did not consider the competition with other charge disordered phases in a fully unbiased calculation as carried out in Refs.. Here a particular feature of the CO $`CE`$-phase structure should be remarked. In the $`x`$-$`y`$ plane, the Mn<sup>3+</sup> and Mn<sup>4+</sup> ions arrange in a checkerboard pattern, which naively seems to arise quite naturally from the presence of a strong long-range Coulomb interaction. However, contrary to this naive expectation, the same CO pattern stacks in experiments along the $`z`$-axis without any change, although the $`t_{2\mathrm{g}}`$ spin direction alternates from plane to plane. In other words, the Mn<sup>3+</sup> and Mn<sup>4+</sup> ions in the $`CE`$-type AFM structure do $`not`$ form a three-dimensional (3D) Wigner-crystal structure, clearly suggesting that the origin of the CO phase in the half-doped manganite $`cannot`$ be caused exclusively by a dominant long-range Coulomb interaction. Other interactions should be important as well. In this work, it is shown that the charge-stacked (CS) CO-state in the $`CE`$-type structure found in experiments can also be obtained in the simple MFA for the purely JT-phononic model. In addition, it is shown that this structure is not easily destroyed by the inclusion of the nearest-neighbor Coulomb interaction. The exchange $`J_{\mathrm{AF}}`$ between the $`t_{2\mathrm{g}}`$ spins plays a key role in the stabilization of the CS structure. The details of the calculation are discussed here. The present semi-analytical results are in excellent agreement with recent Monte Carlo (MC) simulations and provide a simple formalism to rationalize these computational results. On the other hand, it is possible to take another approach to understand the CO state in manganites including the $`CE`$-type structure, by emphasizing the role of the on-site Coulomb interaction. In this direction of study, several results are available in the literature. For instance, Ishihara et al. discussed the spin-charge-orbital ordering from the limit of infinite strong electronic correlation, namely, under the approximation of no double occupancy. Maezono et al. provided information on the overall features of the phase diagram for manganites by using a mean-field calculation. These two mean-field approaches do not reproduce the $`CE`$-state of half-doped manganites. Mizokawa and Fujimori discussed the stabilization of the $`CE`$-type structure by using the model Hartree-Fock approximation. However, to the best of our knowledge, the origin of the CS structure in the $`CE`$-type state has not been clarified based on the Coulombic scenario, although several properties of manganites have been reproduced. In this paper, a possible way to understand the $`CE`$-type state with the CS structure in the purely Coulombic model even without the long-range repulsion is briefly discussed. The present paper contains information about other subjects as well. In spite of the importance of additional interactions besides the long-range Coulombic one to stabilize the proper charge-stacked $`n`$=$`0.5`$ state, as discussed before, it is clear that at least the on-site Coulomb interactions are very strong and their influence should be considered in realistic calculations. Most of the previous analytic and computational works that reported the PS tendency and the CO phase have used electrons interacting among themselves indirectly through their coupling with localized $`t_{2\mathrm{g}}`$ spins or with JT phonons. The JT-phononic model has been extensively studied by the present authors since the explicit inclusion of on-site Coulomb interactions diminish dramatically the feasibility of the computational work. In addition, it has been found that the JT-phononic model explains quite well several experimental results, even if the Coulomb interaction is not included explicitly. Moreover, it should be noted that the energy gain due to the static JT distortion is maximized when only one $`e_\mathrm{g}`$ electron is located at the JT center. Confirming this expectation, the double-occupancy of a given orbital has been found to be negligible in previous investigations at intermediate and large values of the electron-phonon coupling, and in this situation, adding an on-site repulsion would not alter the physics of the state under investigation. Nevertheless, although the statements above are believed to be qualitatively correct, it should be checked explicitly in a more realistic model that indeed the physics found in the extreme case of only phononic interactions survives when both JT-phononic and Coulombic interactions are considered. Thus, another purpose of this paper is to clarify the validity of the JT-phononic only model by showing that the inclusion of the Coulomb interaction does not bring qualitative changes in the conclusions obtained from the purely phononic model. For this purpose, an analytic-numeric combination is employed, namely, the MFA including Coulombic terms is developed in detail and the results are compared with exact diagonalization (ED) data for a small cluster. In addition, the PS tendency is reexamined briefly within the framework of MFA, and some comments on previous results are provided. It is clear that addressing whether purely JT phononic or purely Coulombic approaches lead or not to qualitatively different phase diagrams is an important area of investigation. As a special case of the general goal described in the previous paragraph, the issue of PS in the presence of Coulomb interactions will also be studied in this paper using computational techniques. In fact, it has not been analyzed whether models with two orbitals and Coulomb repulsions, i.e. without JT-phonons, lead to mixed-phase tendencies as pure phononic models do. In this paper it is reported that a study of a simple one-dimensional two-orbital model without phonons but with strong Coulomb interactions, produces PS tendencies similarly as reported in previous investigations by our group. Several features are in qualitative agreement with those observed using JT-phonons, illustrating the similarities between the purely phononic and the purely Coulombic approaches to manganites, and the clear relevance of mixed-phase characteristics for a proper description of these compounds. It is concluded that simple models for manganites, once studied with robust computational techniques, are able to reveal a complex phase diagram with several phases having characteristics quite similar to those observed experimentally. The organization of the paper is as follows. In Sec. II, a theoretical model for manganites is introduced, and several reduced Hamiltonians are discussed in preparation for the subsequent investigations. Section III is devoted to the MFA. The formulation for the MFA is provided and a comparison with the ED result is shown to discuss the validity of the MFA. It is argued that the JT-phononic interaction plays a role more relevant than the Coulombic interaction as suggested from the viewpoint of the continuity of the orbital-ordered (OO) phase. In Sec. IV, as an application of the present MFA, the CO/OO structure in half-doped manganites is studied. Especially, the origin of the $`CE`$-type state with the CS structure is discussed in detail. In Sec. V, the PS tendency is studied in the framework of the MFA for the purely JT-phononic and Coulombic models. In Sec. VI, the two-orbital 1D Hubbard model is analyzed using the density matrix renormalization group (DMRG) technique. Even in the purely Coulombic model, PS tendencies and CO phases are observed. Finally in Sec. VII, the main results of the paper are summarized. The main overall conclusion is that studies based upon purely JT-phononic or Coulombic approaches do $`not`$ differ substantially in their results, and there is no fundamental difference between them. For reasons to be described below, of the two approaches the best appears to be the phononic one due to its ability to select the proper orbital ordering pattern uniquely. PS appears in purely phononic and purely Coulombic approaches as well. In Appendix A, the relation between the PS scenarios for cuprates and manganites is discussed. Throughout this paper, units such that $`\mathrm{}`$=$`k_\mathrm{B}`$=1 are used. ## II Models for Manganites In order to explain the existence of ferromagnetism at low temperatures in the manganites, the DE mechanism has been usually invoked. In this framework, holes can improve their kinetic energy by polarizing the spins in their vicinity. If only one $`e_\mathrm{g}`$ orbital is used, the FM Kondo or one-orbital model is obtained, and several results for this model are already available in the literature. Although the important PS tendencies have been identified in this context, it is clear that the one-orbital model is not sufficient to describe the rich physics of the manganites, where the two active orbitals play a key role and the CO state close to density 0.5 is crucial for the appearance of a huge MR effect at this density. Thus, in this paper, the two-orbital model is discussed to understand the manganite physics, by highlighting the complementary roles of the JT phonon and Coulomb interactions. ### A Hamiltonian Let us consider doubly-degenerated $`e_\mathrm{g}`$ electrons, tightly coupled to localized $`t_{2\mathrm{g}}`$ spins and local distortions of the MnO<sub>6</sub> octahedra. This situation is described by the Hamiltonian $`H`$ composed of five terms $$H=H_{\mathrm{kin}}+H_{\mathrm{Hund}}+H_{\mathrm{AFM}}+H_{\mathrm{el}\mathrm{ph}}+H_{\mathrm{el}\mathrm{el}}.$$ (1) The first term indicates the hopping motion of $`e_\mathrm{g}`$ electrons, given by, $`H_{\mathrm{kin}}={\displaystyle \underset{\mathrm{𝐢𝐚}\gamma \gamma ^{}\sigma }{}}t_{\gamma \gamma ^{}}^𝐚d_{𝐢\gamma \sigma }^{}d_{𝐢+𝐚\gamma ^{}\sigma },`$ (2) where $`d_{𝐢\mathrm{a}\sigma }`$ ($`d_{𝐢\mathrm{b}\sigma }`$) is the annihilation operator for an $`e_\mathrm{g}`$-electron with spin $`\sigma `$ in the $`d_{x^2y^2}`$ ($`d_{3z^2r^2}`$) orbital at site $`𝐢`$, $`𝐚`$ is the vector connecting nearest-neighbor sites, and $`t_{\gamma \gamma ^{}}^𝐚`$ is the nearest-neighbor hopping amplitude between $`\gamma `$\- and $`\gamma ^{}`$-orbitals along the $`𝐚`$-direction. The amplitudes are evaluated from the overlap integral between manganese and oxygen ions, given by $`t_{\mathrm{aa}}^𝐱=\sqrt{3}t_{\mathrm{ab}}^𝐱=\sqrt{3}t_{\mathrm{ba}}^𝐱=3t_{\mathrm{bb}}^𝐱=t,`$ (3) for the $`x`$-direction, $`t_{\mathrm{aa}}^𝐲=\sqrt{3}t_{\mathrm{ab}}^𝐲=\sqrt{3}t_{\mathrm{ba}}^𝐲=3t_{\mathrm{bb}}^𝐲=t,`$ (4) for the $`y`$-direction, and $`t_{\mathrm{bb}}^𝐳=4t/3,t_{\mathrm{aa}}^𝐳=t_{\mathrm{ab}}^𝐳=t_{\mathrm{ba}}^𝐳=0,`$ (5) for the $`z`$-direction. Note that $`t_{\mathrm{aa}}^𝐱`$ (equal to $`t_{\mathrm{aa}}^𝐲`$ by symmetry) is taken as the energy scale $`t`$. The second term is the Hund coupling between the $`e_\mathrm{g}`$ electron spin $`𝐬_𝐢`$ and the localized $`t_{2\mathrm{g}}`$ spin $`𝐒_𝐢`$, given by $$H_{\mathrm{Hund}}=J_\mathrm{H}\underset{𝐢}{}𝐬_𝐢𝐒_𝐣,$$ (6) with $`𝐬_𝐢=_{\gamma \alpha \beta }d_{𝐢\gamma \alpha }^{}𝝈_{\alpha \beta }d_{𝐢\gamma \beta }`$, where $`J_\mathrm{H}`$($`>`$0) is the Hund coupling, and $`𝝈`$=$`(\sigma _x,\sigma _y,\sigma _z)`$ are the Pauli matrices. Note here that the $`t_{2\mathrm{g}}`$ spins are assumed classical and normalized to $`|𝐒_𝐢|`$=1. The third term accounts for the $`G`$-type AFM property of manganites in the fully hole doped limit, given by $$H_{\mathrm{AFM}}=J_{\mathrm{AF}}\underset{𝐢,𝐣}{}𝐒_𝐢𝐒_𝐣,$$ (7) where $`J_{\mathrm{AF}}`$ is the AFM coupling between nearest neighbor $`t_{2\mathrm{g}}`$ spins. The value of $`J_{\mathrm{AF}}`$ will be small in units of $`t`$ to make it compatible with experiments for the fully hole-doped CaMnO<sub>3</sub> compound. In the fourth term, the coupling of $`e_\mathrm{g}`$ electrons to the lattice distortion is considered as $`H_{\mathrm{el}\mathrm{ph}}=g{\displaystyle \underset{𝐢\sigma }{}}[Q_{1𝐢}(d_{𝐢a\sigma }^{}d_{𝐢a\sigma }+d_{𝐢b\sigma }^{}d_{𝐢b\sigma })`$ (8) $`+`$ $`Q_{2𝐢}(d_{𝐢a\sigma }^{}d_{𝐢b\sigma }+d_{𝐢b\sigma }^{}d_{𝐢a\sigma })+Q_{3𝐢}(d_{𝐢a\sigma }^{}d_{𝐢a\sigma }d_{𝐢b\sigma }^{}d_{𝐢b\sigma })]`$ (9) $`+`$ $`(1/2){\displaystyle \underset{𝐢}{}}[k_{\mathrm{br}}Q_{1𝐢}^2+k_{\mathrm{JT}}(Q_{2𝐢}^2+Q_{3𝐢}^2)],`$ (10) where $`g`$ is the coupling constant between $`e_\mathrm{g}`$ electrons and distortions of the MnO<sub>6</sub> octahedron, $`Q_{1𝐢}`$ is the breathing-mode distortion, $`Q_{2𝐢}`$ and $`Q_{3𝐢}`$ are, respectively, the JT distortions for the $`(x^2y^2)`$\- and $`(3z^2r^2)`$-type modes, and $`k_{\mathrm{br}}`$ ($`k_{\mathrm{JT}}`$) is the spring constant for the breathing (JT) mode distortions. Here the distortions are treated adiabatically, but the validity of this treatment should be discussed. In general, the adiabaticity is judged from the ratio $`\eta `$= $`\tau _\mathrm{e}/\tau _\mathrm{i}`$, where $`\tau _\mathrm{e}`$ and $`\tau _\mathrm{i}`$ are characteristic time scales for the motion of electrons and ions, respectively. If $`\eta `$ is much less than unity, the adiabatic approximation holds, since the motion of ions is very slow compared to that of electrons. Due to the relations $`\tau _\mathrm{i}`$$``$$`\omega ^1`$ and $`\tau _\mathrm{e}`$$``$$`t^1`$, where $`\omega `$ is the frequency of the lattice distortion, $`\eta `$$``$$`\omega /t`$ is obtained. In the manganite, $`t`$ is 0.2-0.5 eV, while $`\omega `$ is about $`500`$-$`600`$cm<sup>-1</sup>. Thus, the adiabatic approximation is in principle valid in the study of the JT distortion of manganites. An important energy scale characteristic of $`H_{\mathrm{el}\mathrm{ph}}`$ is the static JT energy, defined by $`E_{\mathrm{JT}}=g^2/(2k_{\mathrm{JT}}),`$ (11) which is naturally obtained by the scaling of $`Q_{\mu 𝐢}=(g/k_{\mathrm{JT}})q_{\mu 𝐢}`$ with $`\mu `$=1,2, and 3, where $`g/k_{\mathrm{JT}}`$ is the typical length scale for the JT distortion. By using $`E_{\mathrm{JT}}`$ and $`t`$, it is convenient to introduce the non-dimensional electron-phonon coupling constant $`\lambda `$ as $`\lambda =\sqrt{2E_{\mathrm{JT}}/t}.`$ (12) The characteristic energy for the breathing-mode distortion is given by $`E_{\mathrm{br}}`$=$`g^2/(2k_{\mathrm{br}})`$=$`E_{\mathrm{JT}}/\beta `$ with $`\beta `$=$`k_{\mathrm{br}}/k_{\mathrm{JT}}`$. The last term indicates the Coulomb interactions between $`e_\mathrm{g}`$ electrons, expressed by $`H_{\mathrm{el}\mathrm{el}}`$ $`=`$ $`U{\displaystyle \underset{𝐢\gamma }{}}\rho _{𝐢\gamma }\rho _{𝐢\gamma }+U^{}{\displaystyle \underset{𝐢\sigma \sigma ^{}}{}}\rho _{𝐢\mathrm{a}\sigma }\rho _{𝐢\mathrm{b}\sigma ^{}}`$ (13) $`+`$ $`J{\displaystyle \underset{𝐢\sigma \sigma ^{}}{}}d_{𝐢\mathrm{a}\sigma }^{}d_{𝐢\mathrm{b}\sigma ^{}}^{}d_{𝐢\mathrm{a}\sigma ^{}}d_{𝐢\mathrm{b}\sigma }+V{\displaystyle \underset{𝐢,𝐣}{}}\rho _𝐢\rho _𝐣,`$ (14) with $`\rho _{𝐢\gamma \sigma }`$= $`d_{𝐢\gamma \sigma }^{}d_{𝐢\gamma \sigma }`$ and $`\rho _𝐢`$=$`_{\gamma \sigma }\rho _{𝐢\gamma \sigma }`$, where $`U`$ is the intra-orbital Coulomb interaction, $`U^{}`$ is the inter-orbital Coulomb interaction, $`J`$ is the inter-orbital exchange interaction, and $`V`$ is the nearest-neighbor Coulomb interaction. Note that the Hund-coupling term $`H_{\mathrm{Hund}}`$ between conduction and localized spins also arises from Coulombic effects, but in this paper, the “Coulomb interaction” refers to the direct electrostatic repulsion between $`e_\mathrm{g}`$ electrons. Note also that the relation $`U`$=$`U^{}`$+$`2J`$ holds in the localized ion system. Although the validity of this relation may not be guaranteed in the actual material, it is assumed to hold in this paper to reduce the parameter space in the calculation. Finally, a comment about the cooperative JT effect which is not included in the phononic models is discussed here. In principle, a Heisenberg-like coupling between nearest-neighbor phonons should also be added to account for the fact that when a given octahedra is elongated in one direction, its neighbors deform in an opposite way. However, this cooperative effect has not been studied in detail using computational techniques and here the focus of the paper will be on results obtained with non-cooperative phonons. The rich phase diagram and good qualitative agreement with experiments justifies a-posteriori this extra approximation. In fact, in the undoped limit, there are no essential differences between the results with and without the cooperative effect. ### B Reduced Hamiltonians The Hamiltonian Eq. (1) is believed to define an appropriate model for manganites, but it is quite difficult to solve it exactly. In order to investigate further the properties of manganites based on $`H`$, some simplifications and approximations are needed. A simplification without the loss of essential physics is to take the widely used limit $`J_\mathrm{H}`$=$`\mathrm{}`$. In such a limit, the $`e_\mathrm{g}`$ electron spin perfectly aligns along the $`t_{2\mathrm{g}}`$ spin direction, reducing the number of degrees of freedom. Moreover, there is a clear advantage that both the intra-orbital on-site term and the exchange term can be neglected in $`H_{\mathrm{el}\mathrm{el}}`$. Thus, the following simplified model is obtained: $`H^{\mathrm{}}={\displaystyle \underset{\mathrm{𝐢𝐚}\gamma \gamma ^{}}{}}𝒮_{𝐢,𝐢+𝐚}t_{\gamma \gamma ^{}}^𝐚c_{𝐢\gamma }^{}c_{𝐢+𝐚\gamma ^{}}+J_{\mathrm{AF}}{\displaystyle \underset{𝐢,𝐣}{}}𝐒_𝐢𝐒_𝐣`$ (15) $`+`$ $`E_{\mathrm{JT}}{\displaystyle \underset{𝐢}{}}[2(q_{1𝐢}n_𝐢+q_{2𝐢}\tau _{x𝐢}+q_{3𝐢}\tau _{z𝐢})+\beta q_{1𝐢}^2+q_{2𝐢}^2+q_{3𝐢}^2]`$ (16) $`+`$ $`U^{}{\displaystyle \underset{𝐢}{}}n_{𝐢\mathrm{a}}n_{𝐢\mathrm{b}}+V{\displaystyle \underset{𝐢,𝐣}{}}n_𝐢n_𝐣,`$ (17) where $`c_{𝐢\gamma }`$ is the spinless $`e_\mathrm{g}`$ electron operator, given by $`c_{𝐢\gamma }`$= $`\mathrm{cos}(\theta _𝐢/2)d_{𝐢\gamma }`$\+ $`\mathrm{sin}(\theta _𝐢/2)e^{i\varphi _𝐢}d_{𝐢\gamma }`$, the angles $`\theta _𝐢`$ and $`\varphi _𝐢`$ define the direction of the classical $`t_{2\mathrm{g}}`$ spin at site $`𝐢`$, $`n_{𝐢\gamma }`$=$`c_{𝐢\gamma }^{}c_{𝐢\gamma }`$, $`n_𝐢`$=$`_\gamma n_{𝐢\gamma }`$, $`\tau _{x𝐢}`$= $`c_{𝐢a}^{}c_{𝐢b}`$+$`c_{𝐢b}^{}c_{𝐢a}`$, and $`\tau _{z𝐢}`$= $`c_{𝐢a}^{}c_{𝐢a}`$$``$$`c_{𝐢b}^{}c_{𝐢b}`$. Here $`𝒮_{𝐢,𝐣}`$ denotes the change of hopping amplitude due to the difference in angles between $`t_{2\mathrm{g}}`$ spins at sites $`𝐢`$ and $`𝐣`$, given by $`𝒮_{𝐢,𝐣}`$ $`=`$ $`\mathrm{cos}(\theta _𝐢/2)\mathrm{cos}(\theta _𝐣/2)`$ (18) $`+`$ $`\mathrm{sin}(\theta _𝐢/2)\mathrm{sin}(\theta _𝐣/2)e^{i(\varphi _𝐢\varphi _𝐣)}.`$ (19) In principle, $`\theta _𝐢`$ and $`\varphi _𝐢`$ should be optimized to provide the lowest-energy state. However, in the following analytic approach, only spin configurations such that nearest neighbor $`t_{2\mathrm{g}}`$ spins are either FM or AFM will be considered. Namely, the spin canting state is excluded from the outset. The validity of this simplification is later checked by comparison with the numerical results obtained from a relaxation technique. Note that this simplification does not restrict us to only fully FM or $`G`$-type AFM states, but a wide variety of other states, such as the $`CE`$-type one, can also be studied. Based on the reduced Hamiltonian $`H^{\mathrm{}}`$, it is possible to consider two scenarios for manganites, as schematically shown in Fig. 1. Temporarily, $`V`$ will not be taken into account to focus on the competition between $`E_{\mathrm{JT}}`$ and $`U^{}`$. One of them is the JT-phonon scenario in which $`U^{}`$ is considered as a correction into the purely JT-phononic model $`H_{\mathrm{JT}}^{\mathrm{}}`$=$`H^{\mathrm{}}`$($`U^{}`$=0,$`V`$=0) taken as starting point. Another is the Coulomb scenario where the JT distortion is a correction into the purely Coulombic Hubbard-like model $`H_\mathrm{C}^{\mathrm{}}`$=$`H^{\mathrm{}}`$($`E_{\mathrm{JT}}`$=0). Of course, if the many-body analysis were very accurate then there should be no difference between the final conclusions obtained from both scenarios when the realistic situation, $`E_{\mathrm{JT}}`$$`>`$0 and $`U^{}`$$`>`$0, is approached. However, if this realistic situation can be continuously obtained from the limiting cases $`U^{}`$=0 or $`E_{\mathrm{JT}}`$=0, then it is enough to consider only the limiting models $`H_{\mathrm{JT}}^{\mathrm{}}`$ or $`H_\mathrm{C}^{\mathrm{}}`$ to grasp the essence of the manganite physics. On the other hand, it may also occur that starting in one of the two extreme cases a singularity exists preventing a smooth continuation into the realistic region of parameters. In this case, just one of the scenarios would be valid. For investigations in this context, the simplified model $`H^{\mathrm{}}`$ certainly provides a good stage to examine the roles of $`U^{}`$ and $`E_{\mathrm{JT}}`$. In the following section, this point will be discussed in detail. Another possible simplification could have been obtained by neglecting the electron-electron interaction in $`H`$ but keeping the Hund coupling finite, leading to the following purely JT-phononic model with active spin degrees of freedom: $$H_{\mathrm{JT}}=H_{\mathrm{kin}}+H_{\mathrm{Hund}}+H_{\mathrm{AFM}}+H_{\mathrm{el}\mathrm{ph}}.$$ (20) Note that this model is reduced to $`H_{\mathrm{JT}}^{\mathrm{}}`$ when the limit $`J_\mathrm{H}`$=$`\mathrm{}`$ is taken. To solve $`H_{\mathrm{JT}}`$, numerical methods such as MC techniques and the relaxation method have been employed in the past. However, here this model will not be discussed explicitly, but the simplified version $`H_{\mathrm{JT}}^{\mathrm{}}`$ will be investigated in the next section. Qualitatively, the negligible values of the probability of double occupancy at large $`\lambda `$ justifies the neglect of $`H_{\mathrm{el}\mathrm{el}}`$. In addition, the results of recent studies addressing the $`A`$-type AFM-phase of the hole-undoped limit using cooperative JT-phonons have been shown to be quite similar to those found with pure Coulombic interactions, the latter treated with the MFA. Nevertheless, in spite of the above discussed indications that the JT and Coulomb formalisms lead to similar physics, it would be important to verify this belief by studying a multi-orbital model with only Coulombic terms, without the extra approximation of using mean-field techniques for its analysis. Of particular relevance is whether PS tendencies and charge ordering appear in this case, as they do in the JT-phononic model. This analysis is particularly important since, as explained before, a mixture of phononic and Coulombic interactions is expected to be needed for a proper quantitative description of manganites. For this purpose, yet another simplified model will be analyzed in this paper: $$H_\mathrm{C}=H_{\mathrm{kin}}+H_{\mathrm{el}\mathrm{el}}.$$ (21) Note that this model cannot be exactly reduced to $`H_\mathrm{C}^{\mathrm{}}`$, since the Hund coupling term between $`e_\mathrm{g}`$ electrons and $`t_{2\mathrm{g}}`$ spins is not explicitly included. The reason for this extra simplification is that the numerical complexity in the analysis of the model is drastically reduced by neglecting the localized $`t_{2\mathrm{g}}`$ spins. In the FM phase, this is an excellent approximation but not necessarily for other magnetic arrangements. Nevertheless the authors believe that it is important to establish with accurate numerical techniques whether the PS tendencies are already present in this simplified two-orbital models with Coulomb interactions, even if not all degrees of freedom are incorporated from the outset. Adding the $`S`$=3/2 quantum localized spins to the problem would considerably increase the size of the Hilbert space of the model, making it intractable with current computational techniques. ## III Mean-field approximation for $`H^{\mathrm{}}`$ Even a simplified model such as $`H^{\mathrm{}}`$ is still difficult to be solved exactly except for some special cases. Thus, in this section, the MFA is developed for $`H^{\mathrm{}}`$ to attempt to grasp its essential physics. Note that even at the mean-field level, due care should be paid to the self-consistent treatment for the lift of the double degeneracy of $`e_g`$ electrons. ### A Formulation First let us rewrite the electron-phonon term by applying a simple standard decoupling such as $`q_{2𝐢}\tau _{x𝐢}q_{2𝐢}\tau _{x𝐢}+q_{2𝐢}\tau _{x𝐢}q_{2𝐢}\tau _{x𝐢}`$, where the bracket means the average value using the mean-field Hamiltonian described below. By minimizing the phonon energy, the local distortion is determined in the MFA as $`q_{1𝐢}`$=$`n_𝐢/\beta `$, $`q_{2𝐢}`$=$`\tau _{x𝐢}`$, and $`q_{3𝐢}`$=$`\tau _{z𝐢}`$. Thus, the electron-phonon term in the MFA is given by $`H_{\mathrm{el}\mathrm{ph}}^{\mathrm{MF}}`$ $`=`$ $`2{\displaystyle \underset{𝐢}{}}[E_{\mathrm{br}}n_𝐢n_𝐢+E_{\mathrm{JT}}(\tau _{x𝐢}\tau _{x𝐢}+\tau _{z𝐢}\tau _{z𝐢})]`$ (22) $`+`$ $`{\displaystyle \underset{𝐢}{}}[E_{\mathrm{br}}n_𝐢^2+E_{\mathrm{JT}}(\tau _{x𝐢}^2+\tau _{z𝐢}^2)].`$ (23) Now let us turn our attention to the electron-electron interaction term. At a first glance, it appears enough to make a similar decoupling procedure for $`H_{\mathrm{el}\mathrm{el}}`$. However, such a decoupling cannot be uniquely carried out, since $`H_{\mathrm{el}\mathrm{el}}`$ is invariant with respect to the choice of $`e_g`$-electron orbitals due to the local SU(2) symmetry in the orbital space. Thus, to obtain the OO state, it is necessary to find the optimal orbital set by determining the relevant $`e_\mathrm{g}`$-electron orbital self-consistently at each site. For this purpose, it is useful to express $`q_{2𝐢}`$ and $`q_{3𝐢}`$ in polar coordinates as $`q_{2𝐢}=q_𝐢\mathrm{sin}\xi _𝐢,q_{3𝐢}=q_𝐢\mathrm{cos}\xi _𝐢.`$ (24) In the MFA, $`q_𝐢`$ and $`\xi _𝐢`$ can be determined as $`q_𝐢=\sqrt{\tau _{x𝐢}^2+\tau _{z𝐢}^2},`$ (25) and $`\xi _𝐢=\pi +\mathrm{tan}^1(\tau _{x𝐢}/\tau _{z𝐢}).`$ (26) By using the phase $`\xi _𝐢`$, $`c_{𝐢\mathrm{a}}`$ and $`c_{𝐢\mathrm{b}}`$ are transformed into $`\stackrel{~}{c}_{𝐢\mathrm{a}}`$ and $`\stackrel{~}{c}_{𝐢\mathrm{b}}`$ as $`\stackrel{~}{c}_{𝐢\mathrm{a}}=e^{i\xi _𝐢/2}[c_{𝐢\mathrm{a}}\mathrm{cos}(\xi _𝐢/2)+c_{𝐢\mathrm{b}}\mathrm{sin}(\xi _𝐢/2)]`$ and $`\stackrel{~}{c}_{𝐢\mathrm{b}}=e^{i\xi _𝐢/2}[c_{𝐢\mathrm{a}}\mathrm{sin}(\xi _𝐢/2)+c_{𝐢\mathrm{b}}\mathrm{cos}(\xi _𝐢/2)]`$. Note that the phase factor is needed for the assurance of the single-valuedness of the basis function, leading to the molecular Aharonov-Bohm effect. It should be noted that the phase $`\xi _𝐢`$ determines the electron orbital set at each site. For instance, at $`\xi _𝐢`$=$`2\pi /3`$, “a” and “b” denote the $`d_{y^2z^2}`$\- and $`d_{3x^2r^2}`$-orbital, respectively. In Table. I, the correspondence between $`\xi _𝐢`$ and the local orbital is summarized for several important values of $`\xi _𝐢`$. Note here that $`d_{3x^2r^2}`$ and $`d_{3y^2r^2}`$ never appear as the local orbital set. In recent publications, those were inadvertently treated as an orthogonal orbital set to reproduce the experimental results, but such a treatment is essentially incorrect, since the orbital ordering is not due to the simple alternation of two arbitrary kinds of orbitals, as shown in the following. | $`\xi _𝐢`$ | a-orbital | b-orbital | | --- | --- | --- | | $`0`$ | $`x^2y^2`$ | $`3z^2r^2`$ | | $`\pi /3`$ | $`3y^2r^2`$ | $`z^2x^2`$ | | $`2\pi /3`$ | $`y^2z^2`$ | $`3x^2r^2`$ | | $`\pi `$ | $`3z^2r^2`$ | $`x^2y^2`$ | | $`4\pi /3`$ | $`z^2x^2`$ | $`3y^2r^2`$ | | $`5\pi /3`$ | $`3x^2r^2`$ | $`y^2z^2`$ | TABLE I. Phase $`\xi _𝐢`$ and the corresponding $`e_\mathrm{g}`$-electron orbitals. By the above transformation, $`H_{\mathrm{el}\mathrm{ph}}^{\mathrm{MF}}`$ and $`H_{\mathrm{el}\mathrm{el}}`$ are, respectively, rewritten as $`H_{\mathrm{el}\mathrm{ph}}^{\mathrm{MF}}`$ $`=`$ $`{\displaystyle \underset{𝐢}{}}\{E_{\mathrm{br}}(2n_𝐢\stackrel{~}{n}_𝐢+n_𝐢^2)`$ (27) $`+`$ $`E_{\mathrm{JT}}[2q_𝐢(\stackrel{~}{n}_{𝐢\mathrm{a}}\stackrel{~}{n}_{𝐢\mathrm{b}})+q_𝐢^2]\},`$ (28) and $`H_{\mathrm{el}\mathrm{el}}=U^{}{\displaystyle \underset{𝐢}{}}\stackrel{~}{n}_{𝐢\mathrm{a}}\stackrel{~}{n}_{𝐢\mathrm{b}}+V{\displaystyle \underset{𝐢,𝐣}{}}\stackrel{~}{n}_𝐢\stackrel{~}{n}_𝐣,`$ (29) where $`\stackrel{~}{n}_{𝐢\gamma }=\stackrel{~}{c}_{𝐢\gamma }^{}\stackrel{~}{c}_{𝐢\gamma }`$ and $`\stackrel{~}{n}_𝐢=\stackrel{~}{n}_{𝐢\mathrm{a}}`$+$`\stackrel{~}{n}_{𝐢\mathrm{b}}`$. Note that $`H_{\mathrm{el}\mathrm{el}}`$ is invariant with respect to the choice of $`\xi _𝐢`$. Now let us apply the decoupling procedure as $`\stackrel{~}{n}_{𝐢\mathrm{a}}\stackrel{~}{n}_{𝐢\mathrm{b}}\stackrel{~}{n}_{𝐢\mathrm{a}}\stackrel{~}{n}_{𝐢\mathrm{b}}+\stackrel{~}{n}_{𝐢\mathrm{a}}\stackrel{~}{n}_{𝐢\mathrm{b}}\stackrel{~}{n}_{𝐢\mathrm{a}}\stackrel{~}{n}_{𝐢\mathrm{b}}`$, by noting the relations $`\stackrel{~}{n}_{𝐢\mathrm{a}}=(n_𝐢q_𝐢)/2`$, $`\stackrel{~}{n}_{𝐢\mathrm{b}}=(n_𝐢+q_𝐢)/2`$, and $`n_𝐢=\stackrel{~}{n}_𝐢`$. Then, the electron-electron interaction term is given in the MFA as $`H_{\mathrm{el}\mathrm{el}}^{\mathrm{MF}}`$ $`=`$ $`(U^{}/4){\displaystyle \underset{𝐢}{}}[2n_𝐢\stackrel{~}{n}_𝐢n_𝐢^2+2q_𝐢(\stackrel{~}{n}_{a𝐢}\stackrel{~}{n}_{b𝐢})+q_𝐢^2]`$ (30) $`+`$ $`V{\displaystyle \underset{\mathrm{𝐢𝐚}}{}}[n_{𝐢+𝐚}\stackrel{~}{n}_𝐢(1/2)n_{𝐢+𝐚}n_𝐢].`$ (31) By combining $`H_{\mathrm{el}\mathrm{ph}}^{\mathrm{MF}}`$ with $`H_{\mathrm{el}\mathrm{el}}^{\mathrm{MF}}`$ and transforming $`\stackrel{~}{c}_{𝐢\mathrm{a}}`$ and $`\stackrel{~}{c}_{𝐢\mathrm{b}}`$ into the original operators as $`c_{𝐢\mathrm{a}}`$ and $`c_{𝐢\mathrm{b}}`$, the mean-field Hamiltonian is finally obtained as $`H_{\mathrm{MF}}^{\mathrm{}}`$ $`=`$ $`{\displaystyle \underset{\mathrm{𝐢𝐚}\gamma \gamma ^{}}{}}t_{\gamma \gamma ^{}}^𝐚c_{𝐢\gamma }^{}c_{𝐢+𝐚\gamma ^{}}+J_{\mathrm{AF}}{\displaystyle \underset{𝐢,𝐣}{}}𝐒_𝐢𝐒_𝐣`$ (32) $`+`$ $`\stackrel{~}{E}_{\mathrm{JT}}{\displaystyle \underset{𝐢}{}}[2(\tau _{x𝐢}\tau _{x𝐢}+\tau _{z𝐢}\tau _{z𝐢})+\tau _{x𝐢}^2+\tau _{z𝐢}^2]`$ (33) $`+`$ $`{\displaystyle \underset{𝐢}{}}[(\stackrel{~}{U}^{}/2)n_𝐢+V{\displaystyle \underset{𝐚}{}}n_{𝐢+𝐚}]n_𝐢`$ (34) $``$ $`{\displaystyle \underset{𝐢}{}}[(\stackrel{~}{U}^{}/4)n_𝐢+(V/2){\displaystyle \underset{𝐚}{}}n_{𝐢+𝐚}]n_𝐢,`$ (35) where the renormalized JT energy is given by $`\stackrel{~}{E}_{\mathrm{JT}}=E_{\mathrm{JT}}+U^{}/4,`$ (36) and the renormalized inter-orbital Coulomb interaction is expressed as $`\stackrel{~}{U}^{}=U^{}4E_{\mathrm{br}}.`$ (37) Physically, the former relation indicates that the JT energy is effectively enhanced by $`U^{}`$. Namely, the strong on-site Coulombic correlation plays the same role as that of the JT phonon, at least at the mean-field level, indicating that it is not necessary to include $`U^{}`$ explicitly in the models, as has been emphasized by the present authors in several publications. The latter equation for $`\stackrel{~}{U}^{}`$ means that the one-site inter-orbital Coulomb interaction is effectively reduced by the breathing-mode phonon, since the optical-mode phonon provides an effective attraction between electrons. The expected positive value of $`\stackrel{~}{U}^{}`$ indicates that $`e_\mathrm{g}`$ electrons dislike double occupancy at the site, since the energy loss is proportional to the average local electron number in the mean-field argument. Thus, to exploit the gain due to the static JT energy and avoid the loss due to the on-site repulsion, an $`e_\mathrm{g}`$ electron will singly occupy the site. If $`\beta `$ is unrealistically small and $`E_{\mathrm{br}}`$ becomes much larger than $`U^{}`$, $`\stackrel{~}{U}^{}`$ can be negative, leading to an effective attraction between $`e_\mathrm{g}`$ electrons, and charge-density-wave (CDW) state or $`s`$-wave superconducting phase will appear. If the dynamical effect is correctly taken into account beyond the adiabatic approximation, the competition among CDW, spin-density-wave, and $`d`$-wave superconducting states will occur. This is an interesting point, but in the actual manganite $`\beta `$ is suggested to be larger than unity. It has been shown that the breathing-mode distortion is suppressed for a reasonable choice of parameters, even if it is included in the calculation. Thus, in the following, $`\beta `$ is taken to be infinity for simplicity, and the breathing mode is dropped from the calculation. ### B JT-phononic vs. Coulombic scenario In order to check the validity of the present mean-field treatment, it is necessary to compare the mean-field result with some exact solutions. For this purpose, numerical techniques are here applied in small-size clusters. For a set of lattice distortions, $`\{q_{1𝐢},q_{2𝐢},q_{3𝐢}\}`$, the ground-state energy is calculated by using the ED method to include exactly the effect of electron correlations. By searching for the minimum energy, the set of optimal distortions $`\{q_{1𝐢}^{\mathrm{opt}},q_{2𝐢}^{\mathrm{opt}},q_{3𝐢}^{\mathrm{opt}}\}`$ is determined. In this exact treatment, the phase to characterize the local orbital set is defined by $`\xi _𝐢`$= $`\mathrm{tan}^1(q_{2𝐢}^{\mathrm{opt}}/q_{3𝐢}^{\mathrm{opt}})`$. In this subsection, a small 4-site 1D chain with the periodic boundary condition is used in order to grasp the most basic aspects of the problem with a minimum requirement of CPU time. The small cluster size and the low dimensionality will be severe tests for the MFA. However, if the validity is verified under such severe conditions, the mean-field result can be more easily accepted in the large cluster limit and in higher dimensions. In order to focus our attention on the roles of JT phonon and on-site correlations, the $`t_{2\mathrm{g}}`$ spin configuration is assumed to be ferromagnetic (thus, $`J_{\mathrm{AF}}`$ will not play an important role) and $`V`$ is set as zero. The effect of $`V`$ will be discussed in the next subsection in the context of the CO state of half-doped manganites. The realistic hopping matrix set is used for the 1D chain in the $`x`$-direction, although the conclusions are independent of the direction of the 1D chain. In Figs. 2(a)-(f), the results for $`n`$=1 and $`E_{\mathrm{JT}}`$=$`t`$ are shown. Figure 2(a) contains the ground-state energy plotted as a function of $`U^{}`$ for a fixed value of $`E_{\mathrm{JT}}`$. In general, the many-body effects cannot be perfectly included within the MFA. However, when $`U^{}`$ is introduced to the 1D chain with $`E_{\mathrm{JT}}`$$``$0, the results becomes exact due to the special properties of one dimension and the use of the realistic hopping amplitude. Namely, the hopping direction is restricted only along one axis, and the orbital set is uniquely determined due to the optimization of the JT distortion. In such an orbital polarized situation, the electron correlation is included exactly even in the mean-field level. Unfortunately, for 2D or 3D FM phases, there is no guarantee that the MFA provides the exact results, as discussed briefly in the final paragraph, but the MFA for the Coulomb interactions is still expected to provide the qualitatively correct tendency due to the general expectation that the MFA becomes valid in higher dimensions. However, in the purely JT-phononic model $`H_{\mathrm{JT}}^{\mathrm{}}`$, a remarkable fact appears. Namely, the MFA is always exact irrespective of the cluster size, electron density, and dimensionality in $`H_{\mathrm{JT}}^{\mathrm{}}`$. This is quite natural, since the static distortion gives only the potential energy for the $`e_\mathrm{g}`$ electrons, and this fermionic sector is essentially a one-body problem, even though the potential should be optimized self-consistently. The lattice distortion is basically determined by the local electron density, and the MFA becomes exact in the static limit for the distortion. For the reasons discussed above the MFA works quite well in the JT scenario, but this fact does not mean that the obtained state is trivial, since the orbital degree of freedom is active and several non-trivial OO phases occur. To visualize this result, the orbital densities, $`\stackrel{~}{n}_{𝐢\mathrm{a}}`$ and $`\stackrel{~}{n}_{𝐢\mathrm{b}}`$, are shown in Fig. 2(b), and the phase $`\xi _𝐢`$ is plotted in Fig. 2(c). It is observed that $`\stackrel{~}{n}_{𝐢\mathrm{a}}`$ becomes very small and $`\stackrel{~}{n}_{𝐢\mathrm{b}}`$ is almost unity for all the sites. For even and odd sites, $`\xi _𝐢`$ is given by $`2\pi /3`$ and $`5\pi /3`$, respectively. Namely, the occupied b-orbital is $`d_{3x^2r^2}`$ at the even sites and $`d_{y^2z^2}`$ at the odd sites, respectively. This is simply the orbital-staggered state in the spin FM phase, which has been observed in the MC analysis. The mechanism of its occurrence is quite simple: Imagine the two adjacent sites in which one $`e_\mathrm{g}`$ electron is present at each site. If the limit $`E_{\mathrm{JT}}`$$``$$`t`$ is considered, the energy gain can be evaluated in second order with respect to $`t`$, and this energy is maximized when the occupied orbital in a certain site is exactly the same as the unoccupied orbital in the neighboring site. Namely, the difference in $`\xi _𝐢`$ between two adjacent sites should be $`\pi `$ at $`n`$=1. Note that the value of $`\xi _𝐢`$ is determined by the hopping direction. Next let us examine the Coulombic scenario, still using a small 4-site cluster to compare mean-field results against exact ones. In Figs. 2(d)-(f), the ground-state energy, orbital densities, and the phase are plotted as a function of $`E_{\mathrm{JT}}`$ for a fixed value of $`U^{}`$=$`5t`$. Again it can be observed that the MFA works quite well, except for the case of $`E_{\mathrm{JT}}`$=0. In this pure Coulombic limit, the JT distortion does not occur, and thus, the results are obtained only by the ED for $`H_\mathrm{C}^{\mathrm{}}`$. The ground state energy obtained in the MFA converges to the exact result in the limit of $`E_{\mathrm{JT}}`$$``$ 0. However, no symbols are shown at $`E_{\mathrm{JT}}`$=0 in Figs. 2(e) and (f), since the orbital densities and the phase could not be fixed at $`E_{\mathrm{JT}}`$=0. This is due to the fact that the energy is invariant for any choice of local orbital, indicating that $`\stackrel{~}{n}_{𝐢\mathrm{a}}`$, $`\stackrel{~}{n}_{𝐢\mathrm{b}}`$, and $`\xi _𝐢`$ cannot be determined at each site for $`E_{\mathrm{JT}}`$=0. In other words, the OO state is not uniquely fixed within the purely Coulombic model in the sense that the a special orbital is not specified at each site. However, the orbital staggered tendency is detected if the orbital correlation as a function of distance is studied, as will be discussed in Sec. VI. Now the roles of the JT-phonon and on-site correlation are discussed. At finite electron-phonon coupling, the optimized orbital is determined, and the MFA provides an essentially exact result for the shape of the orbital. Note that carrying out unbiased MC simulations and using the relaxation technique are still important tasks, since the present MFA works quite well only for a $`fixed`$ $`t_{2\mathrm{g}}`$ spin background. Thus, it is an unavoidable step to check whether the assumed $`t_{2\mathrm{g}}`$ spin pattern is really stable or not with the use of unbiased techniques by optimizing the lattice distortions as well as the $`t_{2\mathrm{g}}`$ spin directions simultaneously. In this sense, the analytic mean-field approach becomes very powerful when it is combined with numerical unbiased techniques. In the purely Coulombic model, however, no special orbital is determined at each site due to the local SU(2) symmetry. Only when the JT distortion is included, the optimal orbital set at each site is determined from the competition between the kinetic energy gain and the potential loss due to the lattice distortion. As a consequence, it appears that the natural starting model to understand the properties of manganites should be the JT-phononic model, and it is enough to include the effect of $`U^{}`$ on $`H_{\mathrm{JT}}^{\mathrm{}}`$. Although the validity of this statement has been shown in a limited situation in this paper, it is believed to be correct in other cases. (As for the metal-insulator transition, a comment will be given in the final paragraph of this subsection.) On the other hand, if the starting model is chosen as the purely Coulombic model, the degeneracy in the orbital space is lifted when the JT distortion is introduced, indicating that the ground-state property abruptly changes due to the inclusion of the JT distortion. However, the on-site correlation is still important to achieve effectively the strong-coupling region, since $`U^{}`$ is renormalized into the JT energy. To check whether the statements in the previous paragraph are correct also in the doped case, the same analysis is carried out for $`n`$=0.5, as shown in Figs. 3(a)-(f). Note again that the MFA results are always exact and the physical quantities are continuous as a function of $`U^{}`$ for non-zero values of $`E_{\mathrm{JT}}`$, but the local orbital set cannot be uniquely determined for $`E_{\mathrm{JT}}`$=0. In addition, the results are independent of $`U^{}`$ in Figs. 3(a)-(c). This is understood as follows: If a unitary transformation is introduced to make the hopping matrix diagonal, the model $`H_\mathrm{C}^{\mathrm{}}`$ in the FM-phase is reduced to $`H_{\mathrm{FK}}`$ $`=`$ $`{\displaystyle \underset{\mathrm{𝐢𝐚}}{}}(4t/3)a_𝐢^{}a_{𝐢+𝐚}+U^{}{\displaystyle \underset{𝐢}{}}a_𝐢^{}a_𝐢b_𝐢^{}b_𝐢,`$ (38) where $`a_𝐢`$ and $`b_𝐢`$ are given by $`a_𝐢`$=$`(\sqrt{3}/2)c_{𝐢a}`$+$`(1/2)c_{𝐢b}`$ and $`b_𝐢`$=$`(1/2)c_{𝐢a}`$+$`(\sqrt{3}/2)c_{𝐢b}`$, respectively. In this reduced model, only the “a”-electrons can hop and they interact with the localized “b”-electrons, which is just the Falicov-Kimball model. Of course, this model is obtained trivially if the 1D chain along the $`z`$-direction is considered. Thus, the ground state at $`n`$=0.5 is obtained by filling the lower “a”-band up to $`n`$=0.5 and $`U^{}`$ does not work at all. When the electron-phonon coupling is switched on, the system becomes insulating, but the effect of $`U^{}`$ is still inactive. Although the perfect agreement of the mean-field results with the exact numbers is due to the special properties of the 1D chain, the JT-phononic model with the realistic hopping appears to contain the important physics of the problem and the electron correlation is not as crucially important as naively expected. As for the orbital densities at $`n`$=0.5, $`\stackrel{~}{n}_{𝐢\mathrm{a}}`$ is always negligible small, but $`\stackrel{~}{n}_{𝐢\mathrm{b}}`$ for the even sites is almost unity, while $`\stackrel{~}{n}_{\mathrm{b}𝐢}`$ for the odd sites is small, clearly indicating the CO tendency. In this case, $`\xi _𝐢`$ is always given by $`2\pi /3`$ irrespective of the site position, denoting the occurrence of the ferro ordering of the $`d_{3x^2r^2}`$ orbitals. This is easily understood by the DE mechanism in the orbital degree of freedom. Namely, the orbital arranges uniformly to improve the kinetic energy of the $`e_\mathrm{g}`$ electrons, just as the spin does in the FM phase of the doped one-orbital model. Finally, a brief comment on the MFA in higher dimensions is provided. For $`U^{}`$=0, the 1D system is insulating if $`E_{\mathrm{JT}}`$ is non-zero, but in higher dimensions, a metal-insulator transition occurs at some finite value of $`E_{\mathrm{JT}}`$. If $`E_{\mathrm{JT}}`$ is larger than this critical value, the ground state wave function will be smoothly changed as $`U^{}`$ increases. However, when $`E_{\mathrm{JT}}`$ is so small that the system is metallic at $`U^{}`$=0, the inclusion of $`U^{}`$ will bring the metal-insulator transition at some value of $`U^{}`$. The MFA cannot predict correctly this metal-insulator transition point, but in this paper, the CO state is mainly discussed. Namely, $`U^{}`$ ($`E_{\mathrm{JT}}`$) is included into the insulating phase originating from $`E_{\mathrm{JT}}`$($`U^{}`$). Although the properties of the metal-insulator transition in the manganese oxide are quite interesting, this type of analysis will be left for the future. ## IV Charge-Orbital Ordering in half-doped manganites In the previous section, the CO-state with a uniform orbital arrangement has been suggested as the ground-state of the 1D chain at $`n`$=0.5. However, in the real materials a more complicated situation occurs. In the narrow-band manganite such as La<sub>0.5</sub>Ca<sub>0.5</sub>MnO<sub>3</sub> and Nd<sub>0.5</sub>Sr<sub>0.5</sub>MnO<sub>3</sub>, the $`CE`$-type AFM phase is stabilized (See Fig. 4(a)). In this structure, the $`t_{2\mathrm{g}}`$ spins form a ferromagnetic array along a zigzag 1D path, the CO state appears with the checkerboard pattern, and the $`(3x^2r^2)`$/$`(3y^2r^2)`$-type orbital ordering is concomitant to this CO state. The origin of this complex spin-charge-orbital structure has been recently clarified on the basis of the topology of the zigzag structure. Along the $`z`$-axis, the $`CE`$-pattern stacks with the same CO/OO structure, but the coupling of $`t_{2\mathrm{g}}`$ spins along the $`z`$-axis is antiferromagnetic. If only the bi-partite charge structure in the $`x`$-$`y`$ plane is considered, naively the nearest-neighbor Coulomb interaction $`V`$ seems to be the key issue in the formation of the CO/OO phase. However, if $`V`$ is strong enough to bring the CO structure in the $`x`$-$`y`$ plane, the 3D Wigner-crystal structure should appear in the cubic lattice, but this is not observed in the real system. Thus, $`V`$ is not the only ingredient needed for the stabilization of the CO structure of half-doped manganites, as already discussed in the Introduction. What other terms in the Hamiltonian may create a charge-stacked phase? In order to clarify this point, in this subsection 4$`\times `$4$`\times `$4 lattices with the periodic boundary condition are studied as a simple representation of half-doped manganites on the basis of $`H^{\mathrm{}}`$. In Fig. 4(b), energies for several magnetic phases are plotted as a function of $`J_{\mathrm{AF}}`$ for $`\lambda =1.6`$. At this stage in the calculations, $`V`$ is set to zero and its effect will be discussed later. The results of Fig. 4(b) are obtained from the JT-phononic model, but as mentioned in the previous section, it can be interpreted that the effect of $`U^{}`$ is included effectively in the electron-phonon coupling. Although the effect of $`U^{}`$ also appears through the non-trivial term $`(\stackrel{~}{U}^{}/2)_𝐢n_𝐢n_𝐢`$, this term essentially indicates the prohibition of double occupancy, and does not lead to qualitative changes in the results obtained from the JT-phononic model, as long as $`\beta `$ is larger than unity and $`\stackrel{~}{U}^{}`$ is positive. Thus, the results below can be interpreted as arising from a purely JT calculation or a mixture of JT and Coulomb interactions within a mean-field technique. The curves shown are the results in the MFA obtained for several fixed $`t_{2\mathrm{g}}`$ spin patterns, while the solid circles are obtained by the optimization of both the local JT distortion and the $`t_{2\mathrm{g}}`$ spin angles simultaneously. The agreement between the analytic and numerical results is excellent, indicating that the MFA works quite well in the JT-phononic model. For $`J_{\mathrm{AF}}`$$``$0.1$`t`$, a metallic 3D FM-phase is stabilized, since this spin pattern optimizes the kinetic energy. In a very narrow region around $`J_{\mathrm{AF}}`$$``$0.1$`t`$, the $`A`$-type AFM phase occurs. This phase is metallic in the present intermediate coupling, but the uniform $`(x^2y^2)`$-type orbital ordering appears. In the wider region 0.1$`t`$$``$$`J_{\mathrm{AF}}`$$``$0.25$`t`$, the $`CE`$-type AFM structure is the ground state. For unrealistic large values of $`J_{\mathrm{AF}}`$, the $`G`$-type AFM phase is stabilized to gain the magnetic energy. Now let us focus our attention on the stabilization of the $`CE`$-type phase. Due to the competition between the kinetic energy of the $`e_\mathrm{g}`$ electrons and the magnetic energy of the $`t_{2\mathrm{g}}`$ spins, the one-dimensional stripe-like AFM configuration occurs, in which arrays of $`t_{2\mathrm{g}}`$ spins order ferromagnetically along some particular 1D paths. Precisely the shape of the 1D FM path, the zigzag path, is a key to understand the $`CE`$-type phase. To see this, let us compare the $`CE`$-state with the $`C`$-type AFM state, which is characterized by the straight-line FM path. Although the energy of the $`C`$-type structure has the same slope as a function of $`J_{\mathrm{AF}}`$ as the $`CE`$-type, it is not the ground-state, as shown in Fig. 4(b). As for the orbital arrangement, the $`(3x^2r^2)`$-type OO occurs in the $`C`$-type, while the $`(3x^2r^2)/(3y^2r^2)`$-type orbital arrangement appears in the $`CE`$-type state. At a first glance, these two OO-states seem to be quite different, but if the concept of the orbital DE mechanism is employed, it is reasonable that there is no essential distinction between them. Namely, in both cases, the $`e_\mathrm{g}`$ electron orbital is always polarized along the hopping direction of the 1D paths. An essential point for the stabilization of the charge-stacked $`CE`$-type AFM phase is the difference of the $`topology`$ of the 1D paths between the straight and zigzag shapes. Mathematically, the topology of the 1D path can be characterized by “the winding number” $`w`$ defined from the Berry phase connection of the $`e_\mathrm{g}`$-electron wave function along the hopping path. As shown in Ref. , $`w`$ is decomposed into two terms as $`w`$=$`w_\mathrm{g}`$+$`w_\mathrm{t}`$. The former, $`w_\mathrm{g}`$, is called the geometric term, which is $`0`$ ($`1`$) corresponding to the ferro- (antiferro-) arrangement in the orbital sector along the 1D path. As discussed above, due to the orbital DE mechanism, $`w_\mathrm{g}`$ is zero in the doped case irrespective of the shape of the hopping path. The difference between the straight and zigzag paths appears in $`w_\mathrm{t}`$, expressed as $`w_\mathrm{t}`$=$`N_\mathrm{v}/2`$, where $`N_\mathrm{v}`$ is the number of vertices appearing in the unit of the 1D chain. Since $`w_\mathrm{t}`$ is determined only by the shape of the 1D path, it is called the topological term. Of course, $`N_\mathrm{v}`$ is zero for the straight path, while $`N_\mathrm{v}`$ is equal to 2 for the zigzag path. Thus, it is obtained that $`w`$=0 and $`1`$ are the values for the $`C`$\- and $`CE`$-type AFM phases, respectively. To understand why the zigzag chain with $`w`$=1 has the lower energy, it is useful to consider the situation $`U^{}`$=$`E_{\mathrm{JT}}`$=0, in which the $`CE`$-type AFM spin structure characterized by the zigzag 1D chain is stabilized in the picture of a band insulator. Due to the periodic change along the zigzag path in the hopping amplitude as { $`\mathrm{}`$, $`t_{\gamma \gamma ^{}}^𝐱`$, $`t_{\gamma \gamma ^{}}^𝐱`$, $`t_{\gamma \gamma ^{}}^𝐲`$, $`t_{\gamma \gamma ^{}}^𝐲`$, $`\mathrm{}`$ }, the system becomes the band insulator. Although the difference in the hopping amplitudes in the $`x`$\- and $`y`$-directions is only a phase factor in $`t_{\mathrm{ab}}^𝐲`$ and $`t_{\mathrm{ab}}^𝐱`$, a bandgap of the order of $`t`$ develops. On the other hand, note that the $`C`$-type AFM states characterized by the straight path is metallic, since there is no change in the hopping amplitude from bond to bond. In Ref. , it has been clearly shown that the $`CE`$-type structure has the largest bandgap among all the possible types of zigzag hopping paths at $`n`$=0.5. Moreover, in Ref. , if the JT energy is switched on in this phase, it has been shown that the $`CE`$-type phase continues to be the ground state and charge-ordering appears with the $`(3x^2r^2)/(3y^2r^2)`$-type orbital-ordering. Thus, based on the band-insulator picture and the spirit of the adiabatic continuation, the $`CE`$-type AFM structure characterized by the 1D zigzag FM chain is the ground-state. Now the physical meaning of the energy difference between $`CE`$\- and $`C`$-type states is discussed using the the concept of the winding number. For this purpose, an analogy with a typical spin problem is quite useful, as suggested by Takada et al. In the spin problem, by classifying the states with the total spin $`S`$ which is a conserved quantity, the exchange energy is defined by the difference between the energies of the singlet ($`S`$=0) and triplet ($`S`$=1) states. In the present problem, $`w`$ is the topological entity and the conserved quantity. Thus, the energy difference between the states with $`w=1`$ and $`w=0`$, $`J_\mathrm{w}`$, is expected to play a similar role as the exchange energy in the spin problem. As for the magnitude of $`J_\mathrm{w}`$, it can be of the order of $`0.1t`$ from the analysis of the two-site problem, although it depends on the value of $`E_{\mathrm{JT}}`$. Another phase competing with the $`CE`$-type state is the shifted $`CE`$-type (s$`CE`$) structure. This is also obtained by the stacking of Fig. 4(a) along the $`z`$-axis, but one-lattice spacing shifted in the $`x`$\- (or $`y`$-)direction. Due to this shift, the number of FM and AFM bonds becomes equal, and the magnetic energy is exactly canceled. In fact, the energy for the s$`CE`$ structure is independent of $`J_{\mathrm{AF}}`$. The $`CE`$-type phase is stabilized against the s$`CE`$-structure by the magnetic energy, as observed in Fig. 4(b), showing the key role played by $`J_{AF}`$ in models for manganites and likely in the real compounds as well. Let us consider now the effect of $`V`$ on the CO states. For this purpose, the mean-field calculation is carried out based on $`H_{\mathrm{MF}}^{\mathrm{}}`$ for a reasonable parameter set such as $`t`$=0.5eV, $`E_{\mathrm{JT}}`$=0.25eV, and $`U^{}`$=5eV. The phase diagram in the $`(J_{\mathrm{AF}},V)`$ plane is shown in Fig. 4(c). From this figure it is clear that the CS structure occurs for the $`CE`$-type AFM phase in a broad region of parameter space, as deduced from the Fourier transform of the charge correlation which has an observed peak at $`(\pi ,\pi ,0)`$ in the $`CE`$-type state, while a peak appears at $`(\pi ,\pi ,\pi )`$ in the s$`CE`$ and $`C`$-type AFM states. A remarkable fact of Fig. 4(c) is that the CS structure is robust against the inclusion of $`V`$. The boundary between the s$`CE`$\- and $`CE`$-phases is qualitatively understood by the balance of the magnetic energy gain and the charge repulsion loss. On the other hand, the phase boundary between the $`CE`$\- and $`C`$-type AFM states is independent of $`J_{\mathrm{AF}}`$, since those two states have the same magnetic energy. As mentioned above, in this case, the energy $`J_\mathrm{w}`$ due to the difference in the topology of 1D path stabilizes the CS structure in spite of the charge repulsion loss. In fact, the phase boundary exists around $`V`$$``$0.3$`t`$, which is the same order as $`J_\mathrm{w}`$. Although it is difficult to know the exact value of $`V`$ in the actual material, if the simple screened Coulomb interaction is estimated using the large dielectric constant of manganites, $`V/t`$ is estimated to be 0.1$``$0.2, i.e., inside the CS region in our phase diagram. Finally, a comment on the stabilization of the $`CE`$-type structure in the purely Coulombic model is provided. As discussed above, in the case of $`U^{}`$=$`E_{\mathrm{JT}}`$=0, the $`CE`$-type AFM spin structure is found to be the ground-state in the band-insulator picture. If $`U^{}`$ is smoothly switched-on in this phase, still keeping $`E_{\mathrm{JT}}`$=0, the $`CE`$-type phase still continues to be the ground state and charge-ordering appears even without the help of $`V`$, since the local charge in the straight segment of the zigzag chain is larger than that at the corner site. The ground state energy for the zigzag chain is lower than that for the straight chain, and its difference is again of the order of 0.1$`t`$. Since this energy difference can compensate the energy loss due to $`V`$ in the CS structure, it would be possible to understand the CS structure even in the purely Coulombic model on the same topological argument as carried out for the JT-phononic model. Thus, it is possible to fill the 3D cubic lattice by the zigzag 1D chains stacked in the $`b`$\- and $`c`$-axis directions, with the same charge ordering but antiparallel $`t_{2g}`$-spin directions across those 1D chains. Note, however, that the energy is invariant for the choice of local $`e_\mathrm{g}`$-electron orbital and the $`(3x^2r^2)/(3y^2r^2)`$-type orbital arrangement $`cannot`$ be specified in the purely Coulombic model. On the other hand, in the pure JT-phononic model, the $`CE`$-type AFM spin structure, the charge-stacked CO state, and the $`(3x^2r^2)/(3y^2r^2)`$-type OO-phase have been fully understood, as explained before in this section. Thus, based on these results, these authors believe that the purely JT-phononic model is more effective than the purely Coulombic model for the theoretical investigation of manganites, although certainly both lead to very similar physics. ## V Phase-Separation Tendency In previous investigations using the unbiased MC simulations, the PS tendency has been clearly established in manganite models. In the simulations, PS appeared both in the one-orbital FM Kondo model and the two-orbital JT-phononic models. PS occurs due to the balance between the kinetic energy of the $`e_\mathrm{g}`$ electrons and the potential energy due to the background, either the $`t_{2\mathrm{g}}`$ spins or the JT distortion. Due to this difference in the origin of the background potential, two-types of PS exist, spin driven and orbital driven. In the following, the 1D case is used for simplicity. In higher dimensions, the situation will be more complicated, but the essential physics is expected to be captured in the 1D case. One type of PS appears in the two-orbital model driven by the spin sector between the spin FM and AFM phases, mainly in the region between $`x`$$``$0.5 and $`x`$=1.0. To improve the kinetic energy, in this doped situation the $`t_{2\mathrm{g}}`$ spins and $`e_\mathrm{g}`$-electron orbital tend to array in a ferro-manner, but in the heavily doped region close to $`x`$=1.0, the $`t_{2\mathrm{g}}`$ spins order antiferromagnetically to gain the magnetic energy which dominates over the kinetic energy. Thus, in this case, the PS between FM/OF- and AFM/OF-states appears, as clearly indicated in the previous MC simulations. Note that the acronym in front of the slash indicates the spin state, while the acronym after the slash denotes the orbital arrangement, e.g., FM/OF indicates the spin FM and orbital ferro state. Note that this PS is possible in the one-orbital model as well, since the orbital degrees of freedom is not active for this case. Another variety of PS is related to the orbital degrees of freedom, which is tightly coupled to the JT distortion, which mainly exists between $`x`$=0 and $`x`$$``$0.5 in the two-orbitals model. In the undoped situation and for reasonable values of $`J_{\mathrm{AF}}`$, the FM/OAF-state is the ground state, as shown in the previous section. Also in the unbiased MC simulation, this phase has been obtained. When holes are doped, the spin structure is still FM, but the orbital arrangement becomes uniform to gain the kinetic energy by polarizing the orbitals along the hopping direction (orbital DE mechanism). Thus, in this case, the PS between FM/OAF- and FM/OF-states appears, as suggested in the previous works for the two-orbital model. In the MC simulation, the grand canonical ensemble has been used and the chemical potential $`\mu `$ was tuned to obtain the target electron density. In such a calculation, the presence of PS was clearly observed by monitoring the density vs. $`\mu `$. There is a range of densities that cannot be stabilized, no matter how carefully $`\mu `$ is fine-tuned. In other words, the plot $`n`$ as a function of $`\mu `$ presents a discontinuity at some particular value of $`\mu `$. This effect occurs for the one- and two-orbital models, the latter with JT-phonons, and in all dimensions of interest. Further evidence of mixed-phase tendencies can be obtained by monitoring the density as a function of MC-time. In a stable regime, the density does not change much between MC configurations, but in the PS region the fluctuations are very strong. In the present MFA, the electron number $`n`$ has been fixed. For a given $`n`$, the optimized structure both for the $`t_{2\mathrm{g}}`$ spins and the lattice distortion are determined by the analytic MFA and the numerical technique. To understand the PS tendency in this formalism, it is necessary to check the stability of the obtained phase by calculating the ground-state energy $`E_0`$ as a function of $`n`$. Namely, if a negative curvature is obtained, i.e., $`^2E_0/n^2`$$`<`$0, such a phase is unstable even if it is the ground state at a fixed $`n`$. If the PS tendency is included in the present model, the FM/OF states between $`x`$=0 and $`x`$=0.5 should be unstable and the mixed phase should have lower energy than the states obtained by the MFA, in order to reproduce the MC results. In order to verify this, the energy difference $`\mathrm{\Delta }E(x)`$ is plotted as a function of $`x`$ for $`E_{\mathrm{JT}}`$=1.2 in a 16-site spin-FM 1D-chain with the realistic hopping along the $`x`$-axis. Here $`\mathrm{\Delta }E(x)`$=$`E_0(x)`$$``$$`ϵx`$ with $`ϵ`$=$`2[E_0(x`$=$`0)`$$``$$`E_0(x`$=$`0.5)]`$. If $`\mathrm{\Delta }E(x)`$ is positive, the homogeneous state is unstable and, instead, the mixed-phase appears. Note that the state at $`x`$=0.5 is stable. The result is shown in Fig. 5(a), in which positive $`\mathrm{\Delta }E(x)`$ is indeed observed between $`x`$=0 and $`x`$=0.5. In Fig. 5(b), the orbital arrangements at $`x`$=0 and $`x`$=0.5 are shown by depicting the shape of the occupied b-orbital in the $`x`$-$`y`$ plane. It should be noted that the size of the orbital is proportional to $`\stackrel{~}{n}_{b𝐢}`$. This result agrees very well with the previous MC calculation, and the PS tendency is believed to be definitely established in the JT-phononic model. Now let us analyze whether it is possible to detect the PS tendency in the purely Coulombic model or not. For this purpose, $`\mathrm{\Delta }E(x)`$ is evaluated using a 16-site spin-FM 1D-chain with the realistic hopping along the $`x`$-axis for $`E_{\mathrm{JT}}=0`$ and $`U^{}=10t`$ by using the mean-field Hamiltonian Eq. (32). As shown in Fig. 5(c), again the positive $`\mathrm{\Delta }E`$ for $`x`$ between 0 and 0.5 can be observed, indicating clearly the PS tendency. If the present mean-field Hamiltonian is accepted, this result is quite natural, since $`U^{}`$ is included effectively in the coupling between the $`e_\mathrm{g}`$ electrons and JT distortion. However, one may consider that this is just an artifact due to the MFA. Thus, in the following section, it is explicitly shown that this PS tendency in the purely Coulombic model is not due to particular properties of the MFA by performing the DMRG calculation in the 1D Hubbard-like model. In short, it is quite interesting to observe that the MFA can properly reproduce the PS tendencies found using other more sophisticated techniques. Although the calculation is fairly simple and handy the obtained result is physically meaningful and reliable. More investigations on the PS tendency by using the MFA in higher dimensions would be certainly interesting since the MC simulations become increasingly difficult as such dimension grows, but this point will be discussed in future publications. ## VI DMRG result for $`H_\mathrm{C}`$ The purpose of this section is to continue the analysis of the reduced Hamiltonians of Section II, this time focusing on the purely Coulombic model $`H_\mathrm{C}`$. The multi-orbital Hubbard model has been addressed before using a variety of approximations, but here care must be taken to select the appropriate technique accurate enough to search for physics similar to the results obtained with JT-phonons. The use of unbiased methods is particularly important for subtle issues such as PS. For this purpose, the model $`H_\mathrm{C}`$ described in Section II has been studied here with the DMRG method, supplemented by the ED technique, keeping the truncation errors around $`10^6`$ by using typically 120 states on intermediate size chains with open boundary conditions. The restriction to work in the 1D system is not severe in view of the results of previous sections and those of Refs., that showed a strong similarity between one, two and three dimensions, at least regarding the rough features of the ground state. Actually, it is important to remark that the phase diagrams described in previous work by our group are mostly based on evidence coming from the short-distance correlations calculated in our study which are not expected to depend strongly on the dimensionality. In 1D systems, the existence of genuine long-range order vs. slow power-law decay of correlations is a subtle issue beyond the goals of the present analysis. In this context, the static observables studied with the DMRG method are the spin structure factor $$S(k)=\frac{1}{L}\underset{j,m}{}𝐬_j𝐬_me^{i(jm)k},$$ (39) the charge structure factor $$N(k)=\frac{1}{L}\underset{j,m}{}\rho _j\rho _me^{i(jm)k},$$ (40) and the orbital structure factor $$T^z(k)=\frac{1}{L}\underset{j,m}{}T_j^zT_m^ze^{i(jm)k},$$ (41) where $`k`$ is momentum, $`j`$ and $`m`$ denote site positions, $`L`$ is the length of the 1D chain, $`T_j^z`$=$`(\rho _{j\mathrm{a}}`$$``$$`\rho _{j\mathrm{b}})/2`$, and $`\rho _{j\gamma }`$=$`_\sigma \rho _{j\gamma \sigma }`$. It is important to remark that the analysis described below has focussed on the analogies with the results found using JT-phonons, especially regarding PS and the existence and properties of the spin-AF orbital-staggered phase at density $`n`$=1. In the several other interesting phases reported in this section, our effort has been limited to the description of their main features regarding their spin, charge and orbital characteristics, postponing for a future publication a more detailed analysis of its origin and possible relevance to experiments. In order to focus on the similarity in the effects of the on-site correlation and the JT phonons, $`V`$ is set to be zero for the time being. The influence of $`V`$ will be discussed later. Note that the energy unit is $`t`$ in this section, as in the previous ones. ### A Unit Hopping Matrix Thus far the realistic hopping matrix with non-zero off-diagonal hopping amplitude has been used. This hopping matrix is quite important to understand the properties of manganites, but the analysis is sometimes complicated due to the non-zero off-diagonal element. In order to analyze the two-orbital model more easily, it is useful to introduce first a simple unit hopping matrix for any direction, given by $`t_{\gamma \gamma ^{}}^𝐚=t\delta _{\gamma \gamma ^{}},`$ (42) where $`\delta _{ij}`$ is the Kronecker’s delta. Besides its simplicity, this hopping amplitude has been used before in the search for the FM state in the multi-orbital Hubbard model with a variety of techniques, and thus, it is meaningful to investigate its properties. In addition, it will be theoretically interesting to compare results using different hopping sets, to study the dependence of the ground-state properties with those amplitudes. It will actually be concluded that the use of realistic hoppings is very important in this context. #### 1 Case of $`n`$=1 Consider first the special case of density $`n`$=1. Using the DMRG and ED methods a large set of couplings $`(U^{},J)`$ have been investigated, but here only the most representative results are presented as examples. In Fig. 6(a), the spin structure factor $`S(k)`$ is shown at fixed $`U^{}`$=$`10`$, for particular values of $`J`$, and three regimes are clearly identified. At small $`J`$ compared with $`U^{}`$, the spin sector has incommensurate characteristics, with a maximum at momentum $`k`$=$`\pi /2`$. As $`J`$ grows, an abrupt transition to the FM state is observed, with $`S(k)`$ now peaked at zero momentum. This last regime occurs in a robust window of $`J`$. At $`J`$$``$15, the spin sector becomes incommensurate again with a peak in $`S(k)`$ at $`k`$=$`\pi /2`$. The charge structure factor $`N(k)`$ is shown in Fig. 6(b) for the same set of parameters. It is observed that at $`J`$=2 and 5 the charge correlations are not enhanced since only a broad peak with low-intensity appears at $`k`$=$`\pi `$. However, for $`J`$=12 and 20 indications of charge-ordering tendencies are found, since now the $`k`$=$`\pi `$ result is clearly enhanced, indicating a structure with an alternation between the even- and odd-sites of the chain. Note that $`J`$=5 and 12 have different characteristics when analyzed using $`N(k)`$, while they are both ferromagnetic according to $`S(k)`$. In Fig. 7(a), the orbital structure factor $`T^z(k)`$ is presented for the same set of couplings as used in Figs. 6(a) and (b). In this case, enhanced orbital correlations exist in the ground state at $`J`$=2 due to the robust values that $`T^z(k)`$ has, particularly at $`k`$=$`\pi `$. This is indicative of a $`staggered`$ arrangement of orbitals, similar to the results discussed before in Sec. III. A qualitatively similar but even more prominent effect occurs at $`J`$=5. At $`J`$=12 and 20, $`T^z(\pi )`$ is relatively small and there is no noticeable indication of enhanced orbital correlations. The transition from the low-$`J`$ regime, exemplified by $`J`$=2, to the intermediate one ($`J`$=5) is abrupt, as already observed in the study of the spin structure factor. Repeating a similar analysis for several other values of $`U^{}`$ and J allowed us to sketch a phase diagram for the two-orbital Hubbard model at density $`n`$=1, which is shown in Fig. 7(b). Since the complexity of the Hamiltonian does not allow us to perform a careful finite-size study to determine the dominant correlations in the bulk ground-state, the phase diagram of Fig. 7(b) should be considered only qualitative. Nevertheless, since for cluster sizes smaller than $`L`$=20 the results were found to be similar, no large size-effects are anticipated. The four identified regions are marked. Note that the emphasis has been given to the determination of the boundary of the spin-ferromagnetic orbital-ordered phase due to its similarities with the analogous phase reported in the MC studies of the two-orbital model with JT phonons, and in the MFA of the previous section. The boundaries of the other regimes are less accurately determined, particularly at small values of $`J`$ and $`U^{}`$. Let us consider the origin of the many phases observed in Fig. 7. First address the CO regime which appears above the line $`J`$=$`U^{}`$. The intuitive explanation for this behavior is simple. The inter-orbital exchange interaction $`J`$ actually provides an attractive interaction to the electrons, which try to doubly populate half the sites of the chains when this interaction dominates. On the other hand, the inter-orbital repulsion $`U^{}`$ prevents double occupancy of two orbitals at the same site. Then, a competition between the two occurs naturally. If $`J`$$`>`$$`U^{}`$, which is an unphysical limit, then charge-ordering occurs with roughly half the sites having two particles and the other half none. To allow for some nonzero electronic kinetic energy, the doubly occupied sites do not cluster together but are spread on the chain. The spin and charge pattern compatible with the results of Figs.6 and 7 in the CO regime has a unit cell of size four lattice spacings, and an arrangement (2u,0,2d,0) where “2u” denotes two spins up, “0” is an empty site, and “2d” are two spins down. Spin staggered patterns appear frequently in the CO states. The region $`U^{}`$$`>`$$`J`$ is physically more interesting, and connected with the results observed for the model with JT-phonons at the same density. The spin FM characteristic of the FM/OAF phase is believed to be caused by the optimization of the kinetic energy by spin alignment. The influence of the attractive coupling $`J`$ is also very important for the stabilization of this phase, since it also favors spin alignment at every site. The orbital-staggered pattern allows for some mobility of the electrons from site to site, while an orbital-uniform arrangement would not allow for that movement at density $`n`$=1, if all spins were aligned and when the unit-matrix hopping is used. The charge is spread uniformly, i.e., no charge-ordering exists in this phase. However, below but close to the line $`J`$=$`U^{}`$, charge correlations are enhanced due to the proximity to the CO-regime. In this subregion, the orbital correlations are suppressed compared with the results observed at the lower $`J`$ end of the FM/OAF phase, where they are maximized. Overall, it is clear that the FM/OAF state has characteristics very $`similar`$ to those observed when JT-phonons are used to mediate the interaction between electrons (See Sec. III). Then, this phase appears prominently both in studies with phonons and Coulombic interactions, and likely it will be stable when a mixture of the two terms is used. As $`J`$ is decreased for fixed $`U^{}`$, the ferromagnetic tendencies are naturally also reduced. The computational study shows that a regime with sharp spin incommensurate characteristics dominates in this region. The orbital order remains staggered and the charge is uniformly spread. This state is not directly related with the goals of the paper, and thus, the discussion of its origin and characteristics is postponed for future work. #### 2 Case of $`n`$$``$1 One of the main goals of the study in this section is the investigation of whether PS tendencies appear in a multi-orbital model having only Coulomb interactions, starting at $`n`$=1 in the FM/OAF regime previously observed with JT-phonons. For this purpose, here the couplings were fixed to $`U^{}`$=30 and $`J`$=22, i.e., inside the FM/OAF phase of Fig. 7(b), and the density was varied using a cluster of 20 sites. As the number of electrons $`N`$ was changed between 6 and 28 (only using an even number), it was observed that the ferromagnetic characteristics persist and $`S(k)`$ continued to be sharply peaked at zero momentum. Regarding charge-ordering, an enhancement in this channel was found at density $`n`$=0.5 as can be observed in Fig. 8(a) where results for $`N`$=8, 10, 16, and 20 are presented. This enhancement shows that at the couplings studied here tendencies toward a CO-state are developed at $`n`$=0.5, in agreement with recent MC studies for cooperative JT-phonons, the analysis of the previous sections with non-cooperative phonons, and with experiments for manganites. The results in the orbital sector are shown in Fig. 8(b). As the density is reduced, or increased, starting at $`n`$=1 the orbital correlations develop incommensurate characteristics, which is a curious effect not observed before to the best of our knowledge. When the density reaches $`n=0.5`$, $`T^z(k)`$ peaks at $`k`$=$`\pi /2`$, effect likely correlated with the precursors of charge-ordering found in $`N(k)`$. Note that with the DMRG and ED techniques, which are setup in the canonical ensemble where the number of particles can be fixed arbitrarily, the stability of the various states with $`N`$-electrons cannot be addressed directly. For this purpose it is necessary to compare the energies of the various ground states and construct a plot of the density $`n`$ vs. $`\mu `$ following steps already described in detail in previous publications, and in the previous section. Figure 9(a) shows that $`all`$ the densities studied here are actually stable in the sense that a finite window of $`\mu `$ exists for all of them where the state under study minimizes the energy. This has to be contrasted with the results found in the MC and mean-field calculations with JT-phonons where states in a finite window of $`n`$ were found to be unstable, i.e. there was no value of $`\mu `$ that render them the global ground-state of the system. Then, it is concluded that the two-orbital Hubbard model with unit-matrix hopping studied here does not phase-separate in spite of having characteristics at $`n`$=1 very similar to those observed in the MC simulations of the JT-model at the same density. Having a FM/OAF state at $`n`$=1 apparently is not sufficient for PS to occur. This issue is conceptually important and will be addressed again in the next subsection when results with a more realistic nondiagonal hopping matrix are analyzed. Due to the stability of the intermediate phases away from $`n`$=1 it is not too surprising to observe a phase diagram at $`n`$=0.5 with similar characteristics to those found in Fig. 7(b). In Fig. 9(b), the unphysical region $`J`$$`>`$$`U^{}`$ has only been analyzed briefly, just sufficiently to confirm that CO characteristics exist there, with relevant momentum $`k`$=$`\pi /2`$. In the regime $`J`$$`<`$$`U^{}`$, the spin FM and incommensurate phases are still stable, and $`T^z(k)`$ presents a broad peak at $`k`$=$`\pi /2`$ compatible with the tendencies toward charge-ordering that appear at $`n`$=0.5. The probable pattern of orbitals here may have orbital “a” below “b” at site i, a mostly empty i+1 site, the reverse orbital pattern at i+2, and another empty site at i+3, with this arrangement repeated in space. ### B Nondiagonal Hopping Matrix The unit hopping matrix used in the previous subsection is a simple and natural choice to gather qualitative information about the two-orbital problem. However, the studies of models designed for manganites need a more complicated hopping matrix, as shown in Sec. III. In this subsection, the realistic hopping amplitudes along the $`y`$-direction are used to contrast the results against those obtained with the unit hopping matrix. The results are actually found to be dramatically different at densities different from unity. The organization of the subsection is similar as the previous one, i.e., the analysis starts with density $`n`$=1, establishing the main features of the phase diagram, and continues with $`n`$$``$1 with emphasis on the possible appearance of PS. #### 1 Case of $`n`$=1 In Fig. 10(a), the spin structure factor $`S(k)`$ is shown at $`U^{}`$=30 for representative values of $`J`$. At $`J`$=10 and smaller, the signal was found to be clearly antiferromagnetic with a sharp peak at $`k`$=$`\pi `$. In the intermediate region, the spin structure is complex with some incommensurate characteristics, as exemplified by the result at $`J`$=12. For larger values of $`J`$, such as 15 and 20, the system becomes ferromagnetic or quasi-ferromagnetic. If $`J`$ is increased further beyond the $`J`$=$`U^{}`$ boundary, spin incommensurate structures have been observed, as in the study leading to Fig. 7(b) (results in this unphysical regime will not be discussed further here). In Fig. 10(b), the charge structure factor $`N(k)`$ is shown at the same couplings used in Fig. 10(a). This quantity only develops some nontrivial structure as the $`J`$=$`U^{}`$ line is approached. In this case indications of charge-ordering at $`k`$=$`\pi `$ are observed. In Fig. 11(a), the results corresponding to the orbital structure factor are shown. At small $`J`$, the order is uniform since $`T^z(k)`$ peaks at $`k`$=0. At $`J`$=12, an incommensurate structure appears in the orbital sector, similar to the results obtained for the unit-matrix hopping away from $`n`$=1, and related with the spin incommensurability observed in $`S(k)`$. For $`J`$=15 and 20, a robust orbital-staggered pattern is reached. With the information obtained at $`U^{}`$=30 and various $`J`$’s, supplemented by results gathered at $`U^{}`$=5, 10 and 20, a rough phase diagram can be constructed, shown in Fig. 11(b). In this case four regimes are identified. At small $`J`$, an AF/OF state appears which does not exist for the unit-matrix hopping. The reason is the following: For the realistic hopping matrix, orbital “a” has more mobility than the other. Then, to improve the kinetic energy the ground-state prefers to have those orbitals as the lowest-energy ones at every site. However, at $`n`$=1, a spin aligned orbital-uniform arrangement would not have any mobility due to the Pauli principle. Then, the optimal situation is achieved with antiferromagnetic order in the spin sector. The presence of this state is the main difference at $`n`$=1 between the results found for the unit hopping matrix shown in Fig. 7(b) and the results of Fig. 11(b). On the other hand, note that the AF/OF order does not take advantage of $`J`$ since the use of the hoppings in this state leads to antiparallel spins at the same site and, thus, to an energy penalization proportional to $`J`$. For this reason, as the coupling $`J`$ grows, a transition is expected to a state in close competition with the AF/OF one, namely, the FM/OAF state that has appeared several times in our investigations. In the latter $`J`$ plays an important role since it favors spins alignment when two electrons visit the same site, leading to a spin ferromagnetic state. To optimize the kinetic energy, the optimal orbital pattern must be staggered, as discussed in the previous section when explaining Fig. 7(b). The interpolation between these two competing states is difficult to predict and the computational work suggests that it proceeds through a complicated mixture of FM- and AF-orbital characteristics that lead to an overall spin and orbital incommensurate pattern. Only further studies incorporating finite-size analysis will clarify if this regime is indeed stable in the bulk limit. #### 2 Case of $`n`$$``$1 It is important to investigate how the phase diagram shown in Fig. 11(b) evolves as a function of density. To establish a connection with the results found in the case of the JT-phonons at large $`\lambda `$, once again the FM/OAF state is here studied in detail since a state with similar characteristics was indeed observed at $`n`$=1 (Sec.III) in the mean-field studies with phonons, and the reduction of the density led to a phase-separated regime in simulations. Would the same occur in the purely Coulombic case with non-diagonal hopping? To gain insight, $`S(k)`$ is shown in Fig. 12(a) at some representative densities using couplings $`U^{}`$=30 and $`J`$=15 that correspond to a point inside the FM/OAF phase of Fig. 11(b). It is interesting to observe an $`abrupt`$ change in $`S(k)`$ when $`n`$ is varied from 1 to 0.9, with an incommensurate structure appearing in the latter. This incommensuration continues up to $`n`$=0.5. The charge structure factor $`N(k)`$ also presents an abrupt change away from $`n`$=1, with tendencies to charge ordering maximized at $`n`$=0.5 (not shown). In addition, $`T^z(k)`$ drastically switches from an orbital-staggered pattern at $`n`$=1, to a uniform arrangement at $`n`$$``$1 as shown in Fig. 12(b). At $`n`$=0.5 and working with $`U^{}`$=10 and 30, it was observed that the spin-incommensurate and orbital-uniform characteristics are independent of $`J`$ to a good approximation, as long as $`U^{}`$$`>`$$`J`$. To understand the curious behavior reported in Figs. 12(a) and (b), the density vs. $`\mu `$ is plotted in Fig. 13(a) to search for the PS tendency. An anomalous behavior is observed at $`n`$=0.9, which has a tiny $`\mu `$-window of stability, and thus, a large compressibility $`\kappa `$ (since $`\kappa `$ is proportional to d$`n`$/d$`\mu `$). This behavior indicates a tendency toward PS in the data, and it may occur that small additions to the two-orbital Hubbard model such as JT-phonons or even the analysis of slightly larger clusters, may render the system truly phase-separated. To confirm the presence of PS precursors, the local density $`n_i`$ is shown in Fig. 13(b) at $`n`$=0.9. Note that the DMRG method works with open boundary conditions, and $`n_i`$ is not necessarily equal to $`n`$ at every site due to the lack of translational invariance. It is clear from this figure that large charge inhomogeneities are present in the ground state, while at $`n`$=1 (not shown) $`n_i`$ is almost uniform. Only the center of the chain has a density equal to the average one (0.9), while the ends of the chain have $`n`$ close to unity, and at other sites near the center the density is smaller than 0.9. The minimum in the local density shown in Fig. 13(b) may correspond to a hole doped into the $`n`$=1 system, that has developed polaronic characteristics. By symmetry, the other hole is on the other half of the chain. In Fig. 13(c), two holes generate two minima in the density (again, with the other two holes on the other chain half). These results show that tendencies toward charge inhomogeneities are present in this system, and precursors of PS are observed possibly in the form of polaronic behavior. The purely Coulombic model appears to be at the verge of phase separation. To search for more clear indications of phase-separation, the Coulombic interaction $`U^{}`$ was reduced to 10, and $`J`$ was fixed to 7. This still corresponds to a point at $`n`$=1 inside the spin-FM orbital-staggered phase. The spin and orbital structure factors for several number of electrons were calculated in this case, and the results (not shown) have clear similarities with those obtained at $`U^{}`$=30. However, in this case now the analysis of the density vs $`\mu `$ reveals strong phase separation characteristics between densities 1.0 and 0.7 (Fig. 14 (a)), qualitatively similar to those reported using JT-phonons. In Figs. 14 (b) and (c), the local density away from $`n`$=1 is shown. As in the case reported in Figs. 13 (b) and (c), strong oscillations reveal clear tendencies to phase-separation. Studies at $`U^{}`$=20 and $`J`$=12 but with the realistic hopping matrix have also been carried out as part of this effort. The results are very similar to those shown in Figs. 13 and 14. Summarizing, either clear phase separation or a strong tendency to such phenomenon exists in the 1D purely Coulombic model as long as the hopping matrix is non-diagonal. ### C Influence of Nearest-Neighbor Repulsion As discussed in Sec. IV, $`V`$ should not be too large in the actual material, since if it were strong, the CS structure would be destroyed. However, it is important to understand whether the results change or not with the inclusion of $`V`$ using the same unbiased technique as in the previous subsection. First let us consider density $`n`$=1, on a chain of 20 sites and $`U^{}`$=30. In this case, few modifications were observed compared with the results obtained for $`V`$=0. Namely, (i) the boundary between the spin-AFM and spin-incommensurate phases at small $`J`$, compared with $`U^{}`$, shifted toward a smaller $`J`$, and (ii) close to $`J`$=$`U^{}`$ the charge correlations were enhanced substantially favoring the staggered pattern of charge between $`n_i`$=0 and 2, which is not penalized by $`V`$. Even at values such as $`J`$=20, i.e. not too close to $`J`$=$`U^{}`$, this enhancement was noticeable. However, since there are no indications of such a pattern in experiments for manganites at $`n`$=1, there is no need to analyze this region in more detail in the present study. At $`n`$=0.5, it is naively expected that the $`V`$-term brings the CO state, although only precursors of CO behavior were observed at $`V`$=0 and this density as discussed in the previous subsection. In Fig. 15(a), $`N(k)`$ is shown at $`U^{}`$=10 and $`J`$=6 both at $`V`$=0 and 4. The region near $`k`$=$`\pi `$ is clearly enhanced by $`V`$ as expected. The real-space density is shown in Fig. 15(b) and a clear charge-staggered pattern is visible. The spin structure factor (not shown) is still peaked at $`k`$=$`\pi /2`$ as at $`V`$=0, compatible with a staggered spin-arrangement involving the occupied (even or odd) sites. The orbital structure factor still has a large uniform component, although it has also developed a broad low-intensity peak at $`k`$=$`\pi `$. Overall the characteristics of the state stabilized by $`V`$ are orbital-uniform, charge-staggered with period of two lattice spacings, and spin-staggered over the occupied sites (period of four lattice spacings). Thus, only when a sufficiently large nearest-neighbor repulsion is included, the $`n`$=0.5 CO-state similar to that found in the JT-phononic model is reproduced in the purely Coulombic model in one dimension. This result seems in contradiction with the discussion around the CS structure observed in the half-doped material, since it was concluded that $`V`$ is not the only origin of the CO-state in the manganite. However, as briefly discussed in Sec. III C, if the zigzag 1D chain, not the straight 1D chain, is used in the calculation, the CO phase can be obtained even in the purely Coulombic model without $`V`$, although the $`(3x^2r^2)/(3y^2r^2)`$-type OO-phase cannot be obtained. From this viewpoint, a DMRG study carried out on the zigzag 1D chain will be interesting, but this is left for the future. The analysis of the two-orbital model that contains only Coulomb interactions was instructive in several respects. For instance, the study has shown that a spin-ferromagnetic orbital-staggered (FM/OAF) phase appears naturally in this context at $`n`$=1 for a variety of hopping amplitudes, result in excellent agreement with those observed for a purely JT model, also in agreement with the previous ED studies of Coulombic models. Then, the FM/OAF phase is a robust feature of models for manganites and approximations that attempt to describe these materials $`cannot`$ neglect orbital-ordering. Another important result of this section has been the observation of PS tendencies with Coulombic-only interactions. This was shown to occur in the form of robust precursors of this tendency in some regions of parameter space. However, to observe PS it seems necessary to use realistic electronic hopping amplitudes in the case of Coulombic models. The presence of PS tendencies both using purely JT and Coulombic interactions confirms the robustness of this feature, and it also shows that mixed-phase tendencies cannot be ignored in theoretical studies of manganites. In addition, the similarities between models with JT-phonons or Coulombic repulsions suggest that the technically much simpler studies that use only phonons to mediate the interaction between electrons are qualitatively correct, and likely capture the physics of more involved models where both interactions are included. ## VII Discussion and Summary In this paper, the two-orbital model with the JT-phononic and/or the Coulombic interactions has been studied using a variety of techniques. Three points have been confirmed in this work. (i) The main properties of manganites can be reproduced successfully by the purely JT-phononic model even if the strong on-site correlation is not included explicitly, since the effect of such correlation can be renormalized into the effective electron-phonon coupling. (ii) In particular, in the mean-field level, the JT-phononic model can successfully reproduce the $`CE`$-type AFM phase with the charge-stacked structure of the 3D cubic lattice at $`n`$=0.5, in excellent agreement with the MC simulations of Yunoki et al. Even if the nearest-neighbor repulsion $`V`$ is introduced, this phase is not easily destroyed due to the key role played by the magnetic energy gain, regulated by $`J_{\mathrm{AF}}`$, in the $`CE`$\- vs. s$`CE`$-type competition and the “topological” energy gain in the $`CE`$\- vs. $`C`$-type competition. Particularly, it is stressed that the topology of the zigzag 1D path is the key issue leading to the stabilization of the CO/OO state in the $`CE`$-type structure. The on-site Coulombic model treated in the mean-field approximation and without JT phonons was also found to lead to charge-stacking due to the influence of $`J_{\mathrm{AF}}`$. (iii) The purely Coulombic model behaves in many respects very similarly to the purely JT-phononic one and, in particular, it presents the phase separation tendency, especially when realistic hoppings are used. Summarizing, both approaches to the problem of manganites, based either on Coulomb repulsions or phonons, share common tendencies. This conclusion is in agreement with the recent observation that the percolative character of transitions in manganites and its large magnetoresistance effect arise from the competition between metallic and insulating phases in the presence of disorder, independently on whether these phases are mainly generated by Coulombic or JT interactions. Our results have provided robust arguments suggesting that perceiving the “Coulombic” and “JT-phononic” approaches to manganites as qualitatively different ways to carry out theoretical calculations is likely incorrect. ###### Acknowledgements. The authors thank S. Yunoki, A. Moreo, and S. Kivelson for useful conversations. T.H. has been supported by the Ministry of Education, Science, Sports, and Culture of Japan during his stay in the National High Magnetic Field Laboratory, Florida State University. A.L.M. acknowledges the financial support from Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP-Brazil). E.D. is supported by grant NSF-DMR-9814350. ## A Relation with Phase-Separation Theories for Cuprates Since a substantial portion of the paper is devoted to the issue of phase separation in models with Coulomb interactions, in this Appendix extra comments are provided regarding PS in models for high-temperature superconductors (HTSC), issue which has been under discussion for almost a decade.. In particular, it is important to clarify the relation between HTSC phase separation and the PS phenomenon discussed here for manganites. In fact, some authors strongly believe that any doped correlated insulator should have PS, and in this respect the results for manganite models would be a mere particular case of a more general framework. However, it has to be discussed in detail to what extend there is convincing theoretical evidence that indeed any doped insulator phase separates. If this is not clear, the relation between PS in cuprates and manganites weakens considerably. The discussion in this context also has to involve two other important aspects of the problem, namely the microscopic origin of PS, and its phenomenological consequences, particularly when extended Coulomb interactions are included. There are several differences between the PS phenomena proposed for manganites and cuprates. (i) The properties of the two competing phases are not the same. In manganites, an undoped AFM-phase and a hole-doped FM-state are involved, while in cuprates it is an undoped AFM-state and a hole-doped paramagnet or superconductor. In models for HTSC, ferromagnetism does not play an active role at realistic $`J/t`$ couplings, while it is crucial in manganites. Nevertheless in Ref., PS between a G-type antiferromagnet and a ferromagnet was also discussed, within the framework of the very small $`J/t`$ limit of the $`t`$-$`J`$ model and assuming a fully saturated FM-state. The argument in Ref. can be applied directly to the case of the one-orbital model for manganites to justify the presence of PS in this context. However, note that the AFM-state that is doped in 3D real manganites has staggered spin order only in $`one`$ direction, while it is ferromagnetic in the other two ($`A`$-type AFM order). In fact, PS can occur in two-orbitals models for manganites in 1D and 2D without actually involving the AFM-state, but only having two competing FM-states, as discussed in the main text. Here the orbital degree of freedom and JT phonons play the key role needed for PS, while they are not important in the cuprates. (ii) PS in models for manganites occur even in one-dimension in regions of couplings that are realistic, as described in this paper and previous ones, unlike $`t`$-$`J`$ model results in the same dimension where PS only happens at large values of $`J/t`$, and it does not occur at all in the one-band Hubbard model. (iii) The study of PS in models analyzed in the infinite dimensional limit lead to different conclusions between cuprates and manganites, although this issue is still controversial. Fair is to say that the case of the $`D`$=$`\mathrm{}`$ Hubbard model is subtle and deserves a special discussion ($`D`$ is dimensionality). In principle at $`D`$=$`\mathrm{}`$ there are no indications of PS in the one-band Hubbard model. However, it is not clear if this can be taken as a counterexample of PS in a Mott insulator, as recently remarked in Ref.. In most $`D`$=$`\mathrm{}`$ investigations, the ratio $`J`$/$`t`$ scales to zero like 1/$`\sqrt{D}`$ and as a consequence the physics of PS is difficult to address. In addition, when non-bipartite lattices are used, the AFM order is frustrated. Then, more work is needed in large dimension to clarify the effect of doping on correlated insulators. However, note that in calculations for the one-orbital Kondo model for manganites PS has been clearly observed even at $`D`$=$`\mathrm{}`$ (see Ref.). This suggest that important qualitative differences may exist between the PS phenomena in models for cuprates and manganites. Let us elaborate more on stripes in manganites and its topological vs. non-topological character. While stripe order has been claimed to exist in manganites, note that the regime of hole density in which this observation was made ($`x`$=2/3) is far from the low hole concentration were the discussion of stripes in the cuprates occurs. These authors are not aware of experiments reporting stripes at small $`x`$ in manganites, although charge inhomogeneities of various forms appear in several cases in this regime. In addition, it is unclear whether the bi-stripe structure found by Mori et al. has any topological characteristics. Recent calculations by Hotta et al. have shown that the manganite bi-stripes are better interpreted as arising from zigzag conducting chains running $`perpendicular`$ to the charge stripes once electron-phonon JT-couplings are switched-on. For this reason, it is premature to establish connections between stripes in cuprates and manganites. Regarding the issue of whether any doped correlated insulator produces PS, the following is our understanding of the current theoretical literature for the cuprates. The most clear manifestation of PS appears in the $`t`$-$`J`$ model at large $`J/t`$, where the tendency to form pairs of holes to minimize the number of broken AFM bonds is so intense that clusters of holes are formed instead of individual pairs. Here the attractive potential energy among hole carriers originating in the AFM background dominates over the kinetic energy, that tends to spread particles apart. As $`J/t`$ is reduced it has been a matter of much controversy whether the PS effect survives in the realistic small $`J/t`$ regime of the $`t`$-$`J`$ model. While ED of small clusters, MC simulations, and high-temperature expansions suggested that at $`J/t`$ of order unity the effect would disappear, other arguments and further computationalwork opened the possibility for PS to exist at all values of $`J/t`$. Very recent DMRG studies for ladders of increasing number of legs show that previous calculations may not have been sufficiently accurate and the new results suggest that PS indeed only occurs at intermediate and large values of $`J/t`$ in the 2D $`t`$-$`J`$ model. This same conclusion becomes more clear once extra hopping amplitudes are added to the model. In this case the substantial hole mobility induced by the extra hoppings shifts the PS regime to values of $`J/t`$ larger than in the pure $`t`$-$`J`$ model case. Then, the proposal that any correlated insulator when hole doped should become phase separated is still not confirmed using unbiased techniques in simple models for cuprates. Regarding the phenomenological aspects of the PS regime, the issue of microscopic PS that may appear in manganites has certainly been discussed before by Emery and Kivelson for cuprates. In this context the Coulomb interactions break into small pieces the macroscopic clusters of the two phases in competition, since they have different electronic densities. Stripe patterns emerged from calculations carried out mainly close to the atomic limit, where the attraction leading to PS plus the Coulomb repulsion are in competition. In the context of Nuclear Physics similar patterns have also been discussed. In addition, Nagaev studied the formation of finite size clusters of one phase embedded into the other, mainly for antiferromagnetic semiconductors. Then, the simple picture of a $`stable`$ state formed by small clusters of the competing phases, somewhat similar to the CDW pattern obtained in the 1D calculations with nearest-neighbor repulsions, is certainly common to manganites and cuprates, and it has been described in the context of the Frustrated PS scenario for HTSC. In short, the discussion presented in this Appendix suggests that the PS phenomena in models for manganites and cuprates, while sharing the general common ingredients of any PS regime, are different in origin and they must be considered separately in their study.
warning/0003/cs0003053.html
ar5iv
text
# Security of the Cao-Li Public Key Cryptosystem ## 1. Description of the Cryptosystem The Cao-Li public key cryptosystem was first proposed in . It encrypts messages using a bilinear form that is chosen to permit easy decryption by the Chinese remainder theorem. Public key cryptosystems that are designed along this line are not uncommon in the Chinese cryptographic literature. However, as most of the original papers were published in Chinese, they remained relatively obscure until a few of them were described in (in English) recently. Our description below is based on the latter reference. Let $`p_1,\mathrm{},p_n`$ be $`n`$ distinct primes where $`p_i3(mod4)`$. For $`i=1,\mathrm{},n`$, define $$m_i:=\frac{1}{p_i}\left(\underset{j=1}{\overset{n}{}}p_j\right)\text{.}$$ Compute for each $`m_i`$, an integer $`m_i^{}`$ that satisfies $`m_i^{}m_i1(modp_i)`$ and $`0<m_i^{}<p_i`$. We define positive integers $$\lambda _i:=m_i^{}m_i$$ for $`i=1,\mathrm{},n`$ and the diagonal matrix $$\mathrm{\Lambda }:=diag[\lambda _1,\mathrm{},\lambda _n].$$ Note that (1) $$\lambda _i\delta _{ij}(modp_j)$$ where $`\delta _{ij}`$ is $`1`$ if $`i=j`$ and $`0`$ otherwise. We choose another two invertible $`n\times n`$ lower-triangular matrices $`P_1`$ and $`P_2`$ with non-negative integer entries that are bounded by (2) $$\beta :=\underset{1in}{\mathrm{min}}\sqrt{\frac{p_i}{i(i+1)d}}$$ where $`d1`$ is a chosen positive integer. The secret key comprises the two matrices $`P_1,P_2`$ and the primes $`p_i`$, $`i=1,\mathrm{},n`$. The public key is the $`n\times n`$ symmetric matrix $`B`$ given by $$B:=P_2^TP_1^T\mathrm{\Lambda }P_1P_2\text{.}$$ Let the message block be $`𝐱=(x_1,\mathrm{},x_n)`$ where $`0x_id`$. The ciphertext $`y`$ is computed as $$y=𝐱B𝐱^T\text{.}$$ If we let $`𝐳:=𝐱P_2^TP_1^T`$, then $$y=𝐳\mathrm{\Lambda }𝐳^T=\lambda _1z_1^2+\mathrm{}+\lambda _nz_n^2\text{.}$$ From (1), we have (3) $$z_k^2y(modp_k)\text{.}$$ Keeping in mind that $`P_1^T`$ and $`P_2^T`$ are upper-triangular and their entries are non-negative and bounded by $`\beta `$, we have, from (2) and $`0x_id`$, that (4) $$0z_k\underset{i=1}{\overset{k}{}}\underset{j=i}{\overset{k}{}}d\beta ^2=d\beta ^2\frac{k(k+1)}{2}<\frac{p_k}{2}\text{.}$$ We can carry out decryption as follows. For each $`k=1,\mathrm{},n`$, compute the unique $`z_k`$ satisfying (3) and (4). The message can then be recovered by (5) $$𝐱=𝐳\left(P_2^TP_1^T\right)^1\text{.}$$ Note that since $`p_k3(mod4)`$, effective algorithms for computing square roots $`(modp_k)`$ exist (see ). ## 2. Key Recovery We will first recover $`\mathrm{\Lambda }`$ from $`B`$. Let $`P_1P_2=:P=(p_{ij})_{1i,jn}`$. Then $`P`$ is an invertible lower-triangular matrix with non-negative integral entries by the same properties of $`P_1`$ and $`P_2`$. Since $`P`$ is invertible and has non-negative integral entries, we have $`detP=1`$. Moreover, we also have $`detP=p_{11}\times \mathrm{}\times p_{nn}`$ since $`P`$ is triangular. As all the $`p_{ii}`$’s are non-negative, it then follows that $`p_{ii}=1`$ for $`i=1,\mathrm{},n`$. $`\mathrm{\Lambda }`$ and $`P`$ can be recovered from $`B`$ using an algorithm that is very similar to the algorithm for LU-decomposition of a matrix (the difference being that row reduction is done starting from the bottom rows). Denote the $`i`$th row of $`B`$ by $`𝐛_i=(b_{i1},\mathrm{},b_{in})`$, $`i=1,\mathrm{},n`$. We know immediately that $`b_{nn}=\lambda _n`$. The following shows that Algorithm A indeed yields the required output. Let the $`i`$th row of $`P`$ be $`𝐩_i`$, $`i=1,\mathrm{},n`$. Since $`p_{ji}=0`$ if $`j<i`$ and $`p_{ii}=1`$, we may write $`𝐛_i=\lambda _i𝐩_i+_{j=i+1}^n\lambda _jp_{ji}𝐩_j`$. For each $`i=n1,n2,\mathrm{},1`$, the inner loop of Step 1 effectively does $$𝐛_i𝐛_i\underset{j=i+1}{\overset{n}{}}\frac{b_{ji}}{b_{jj}}𝐛_j\text{.}$$ We shall show inductively that $`𝐛_i`$ is reduced to $`\lambda _i𝐩_i`$ at stage $`i`$: clearly $`𝐛_n=\lambda _n𝐩_n`$; suppose $`𝐛_i`$ is reduced to $`\lambda _i𝐩_i`$ at stage $`i=n1,\mathrm{},nk`$, then at stage $`nk1`$, $`𝐛_{nk1}`$ $`𝐛_{nk1}{\displaystyle \underset{j=nk}{\overset{n}{}}}{\displaystyle \frac{b_{ji}}{b_{jj}}}𝐛_j`$ $`=𝐛_{nk1}{\displaystyle \underset{j=nk}{\overset{n}{}}}\lambda _jp_{ji}𝐩_j`$ $`=\lambda _{nk1}𝐩_{nk1}\text{.}`$ Hence Step 1 reduces $`B=(𝐛_1,\mathrm{},𝐛_n)^T`$ to $`(\lambda _n𝐩_n,\mathrm{},\lambda _n𝐩_n)^T=\mathrm{\Lambda }P`$. Since the diagonal entries of $`P`$ are all $`1`$’s, the diagonal entries of $`\mathrm{\Lambda }P`$ are the required $`\lambda _i`$’s. Consequently, $`P`$ can be recovered by dividing each row by its corresponding diagonal entry. We can now recover the moduli $`p_1,\mathrm{},p_n`$ from $`\lambda _1,\mathrm{},\lambda _n`$. From (1), we see that for a fixed $`i`$, $`p_i\lambda _j`$ for all $`ji`$ and $`p_i\lambda _i1`$. So $$p_id_i:=\mathrm{gcd}(\lambda _1,\mathrm{},\lambda _{i1},\lambda _i1,\lambda _{i+1},\mathrm{},\lambda _n)\text{.}$$ It could of course happen that $`d_ip_i`$ for some $`i`$. So this process only partially recovers the $`p_i`$’s. However our computer simulations (using C++ with LiDIA) show that instances where $`d_ip_i`$ are rare. We shall give some heuristics to substantiate this claim. For $`d_i=p_i`$, it is sufficient that $`\mathrm{gcd}(m_1^{},\mathrm{},m_{i1}^{},m_{i+1}^{},\mathrm{},m_n^{})=1`$. From , we have $$\mathrm{\#}\left\{(a_1,\mathrm{},a_k)^k\right|\mathrm{gcd}(a_1,\mathrm{},a_k)=1,\text{ all }a_iN\}$$ $$=\{\begin{array}{cc}N^k/\zeta (k)+O(N^{k1})\hfill & \text{if }k>2,\hfill \\ 6N^2/\pi ^2+O(N\mathrm{log}N)\hfill & \text{if }k=2.\hfill \end{array}$$ where $`\zeta (s)=_{i=1}^{\mathrm{}}i^s`$ is the Riemann zeta function. Assuming that each $`m_i^{}`$ is randomly distributed in $`\{1,\mathrm{},N\}`$ where $`N:=\mathrm{max}\{p_1,\mathrm{},p_n\}`$, the probability that $`\mathrm{gcd}(m_1^{},\mathrm{},m_{i1}^{},m_{i+1}^{},\mathrm{},m_n^{})=1`$ is then at least $`\zeta (n1)6/\pi ^20.60`$ when $`N`$ is large enough. So we can expect to recover more than half of the $`p_i`$’s. In fact our simulations show that we almost always have $`d_i=p_i`$ and many of the rare exceptions are of the form $`d_i=2p_i`$ where $`p_i`$ can also be recovered easily. ## 3. Conclusion Note that Algorithm A is essentially LU-decomposition and the $`d_i`$’s can be computed using the Euclidean algorithm. Since these two methods can be carried out efficiently, we can easily recover $`P`$ and most of the $`p_i`$’s. It then follows that the Cao-Li cryptosystem is insecure and thus should not be used.
warning/0003/quant-ph0003011.html
ar5iv
text
# References EXTENDED COHERENT STATES AND MODIFIED PERTURBATION THEORY G.M.Filippov Chuvash State University, Cheboksary, Russia E-mail: gennadiy@chuvsu.ru ## Abstract An extended coherent state (ECS) for describing a system of two interacting quantum objects is considered. A modified perturbation theory based on using the ECSs is formulated. PACS numbers: 03.65.Fd, 11.15.Bt, 11.15.Tk 1. Extended coherent states Coherent states were constructed first by Schrödinger and in the last 40 years of the 20th century were widely used in different problems of quantum physics . There are many modifications of coherent states. Recall, for example, the spin coherent states introduced in . A general algebraic approach in the coherent state theory was developed in . The coherent states for a particle on a sphere were applied in to describe the rotator time evolution. Here we propose one more generalization of the theory by introducing the extended coherent state (ECS). Consider a system of an oscillator and a free spinless particle posessing a momentum $`𝐤_0`$. Let $`\widehat{b}^{}`$ and $`\widehat{b}`$ be the ladder operators for the oscillator. Introduce the creation $`\widehat{a}^{}`$ and annihilation $`\widehat{a}`$ operators of Bose type to describe a possible change in the particle’s state (note that the further consideration may be applied just as well to a Fermi particle). Input the operator $$\widehat{Q}=\underset{𝐪}{}h_𝐪\widehat{\rho }_𝐪,$$ (1) where $$\widehat{\rho }_𝐪=\underset{𝐤}{}\widehat{a}_𝐤^{}\widehat{a}_{𝐤+𝐪}$$ is the Fourier component of the density operator and $`h_𝐪`$ \- are coefficients depending on momentum $`𝐪`$. We can construct another linear combination $`\widehat{Q}^{}`$ of operators $`\widehat{\rho }_𝐪`$ with the help of any other set of coefficients $`h_𝐪^{}`$. All these combinations are commutative $$[\widehat{Q},\widehat{Q}^{}]_{}=0$$ because the commutation rule $$[\widehat{\rho }_𝐪,\widehat{\rho }_𝐪^{}]_{}=0$$ (2) is fullfilled for all $`𝐪`$ and $`𝐪^{}`$. Input a vector of state $`|0,𝐤_0)`$, where the first argument $`(0)`$ denotes a ground state of the oscillator and the second one $`(𝐤_0)`$ describes a state of the particle. Define the vector $$|h,𝐤_0>=\mathrm{exp}(\frac{1}{2}\widehat{Q}^{}\widehat{Q})\underset{n=0}{\overset{\mathrm{}}{}}\frac{1}{n!}\left(\widehat{Q}\widehat{b}^{}\right)^n|0,𝐤_0)$$ (3) as an ESC (here we briefly denote by $`h`$ the whole set of coefficients $`h_𝐪`$). Obviously, the vector (3) coincides with the ordinary Schrödinger coherent state (SCS), when one replaces all particle’s operators by their classical equivalents. The ECS describes some state of a system of two interacting quantum objects — the particle and the oscillator. By this circumstance the ECS sufficiently differs from the SCS. We outline the following general properties of the ECS: (1) The ECS is not the eigenvector for $`\widehat{b}`$, but $$\widehat{b}|h,𝐤_0>=\widehat{Q}|h,𝐤_\mathrm{𝟎}>.$$ (4) (2) The operators $`\widehat{\rho }_𝐪`$ only change momenta for all the one-particle states. Hence, the following relations are fullfilled: $$\widehat{\rho }_𝐪|h,𝐤_0>=|h,𝐤_0𝐪>$$ (5) $$\widehat{\rho }_𝐪^{}\widehat{\rho }_𝐪|h,𝐤_0>=|h,𝐤_0>.$$ (6) (3) There is the following representation: $$|h,𝐤_0>=\mathrm{exp}[\widehat{Q}\widehat{b}^{}\widehat{Q}^{}\widehat{b}]|0,𝐤_0)$$ (7) which is equivalent to the relevant representation of the SCS. (4) If $`h_𝐪=g\mathrm{\Delta }\left(𝐪𝐪_0\right)`$ we easily have $$|h,𝐤_0>=\mathrm{exp}\left[\frac{\left|g\right|^2}{2}\right]\underset{n=0}{\overset{\mathrm{}}{}}\frac{g^n}{n!}\left(\widehat{\rho }_{𝐪_0}\widehat{b}^{}\right)^n|0,𝐤_0)$$ (8) and therefore, $$<h,𝐤_0|h^{},𝐤_0^{}>=\mathrm{exp}\left[\frac{1}{2}\left(\left|g\right|^2+\left|g^{}\right|^22g^{}g^{}\right)\right]\mathrm{\Delta }\left(𝐤_0𝐤_0^{}\right).$$ (9) (5) The total amount of ECS are more than sufficient to define the Hilbert space. Following to Klauder (see also ) we can introduce the development of the unity operator $$\widehat{\mathrm{I}}=\underset{𝐤}{}\frac{1}{\pi }d^2z\widehat{Q}|zh,𝐤><zh,𝐤|\widehat{Q}^{}$$ (10) where $`z`$ \- is the complex variable, $`d^2z=d\left[\mathrm{Re}\left(z\right)\right]d\left[\mathrm{Im}\left(z\right)\right]`$. To prove the last equation one may use the integral $$d^2z\left(z^{}\right)^nz^m\mathrm{exp}\left[\left|z\right|^2\widehat{Q}^{}\widehat{Q}\right]\widehat{Q}^{m+1}\left(\widehat{Q}^{}\right)^{n+1}=\pi n!\delta _{nm}.$$ (6) There is the following useful sum rule: $$\underset{𝐤}{}e^{i\mathrm{𝐬𝐤}}\widehat{a}_𝐤|h,𝐤_\mathrm{𝟎}>=e^{i\mathrm{𝐬𝐤}_0}|\alpha )\left|\mathrm{vac}_p\right)$$ (11) where the right-hand side contains a direct product of the SCS for the oscillator $$\left|\alpha \right)=\mathrm{exp}[\frac{1}{2}|\alpha |^2]\underset{0}{\overset{\mathrm{}}{}}\frac{\alpha ^n}{n!}\left(b^{}\right)^n\left|0\right)$$ and a vacuum state of the particle $`\left|\mathrm{vac}_p\right)`$. Here the quantity $`\alpha `$ is given by the formula $$\alpha =\underset{𝐪}{}h_𝐪e^{is𝐪}.$$ To prove the property (6) one should keep in mind the relation: $$\underset{𝐤}{}\widehat{a}_𝐤e^{i\mathrm{𝐤𝐱}}\widehat{\rho }_{𝐪_1}\widehat{\rho }_{𝐪_2}\mathrm{}|0,𝐤_0)=e^{i𝐤_0𝐱}e^{i𝐪_1𝐱}e^{i𝐪_2𝐱}\mathrm{}\left|0\right)\left|vac_p\right)$$ (12) where $`\left|0\right)`$ is the vector of the ground state of the oscillator. 2. Modified perturbation theory ECSs, first introduced in 1983<sup>1</sup><sup>1</sup>1 Extended coherent states were first denoted as ’double coherent’ states or ’modified coherent’ states. arise, for example, in a problem of interaction between a moving particle and an oscillator. The proper Hamiltonian can be represented in the following general form $$\widehat{H}_{int}=\widehat{b}^{}\underset{𝐪}{}g_𝐪\widehat{\rho }_𝐪+\widehat{b}\underset{𝐪}{}g_𝐪^{}\widehat{\rho }_𝐪^{}$$ (13) where $`g_𝐪`$ \- is a coupling function. Since $`\widehat{\rho }_𝐪^{}=\widehat{\rho }_𝐪`$, it should be $`g_𝐪=g_𝐪^{}`$. In most applications the Hamiltonian (13) within the interaction picture depends on time via the density operators $`\widehat{\rho }(t)`$. In these cases we cann’t apply ECS without some modification of the theory. Indeed, instead of relations (2) we have $$[\widehat{\rho }_𝐪\left(t\right),\widehat{\rho }_𝐪^{}\left(t^{}\right)]_{}=\underset{𝐤}{}\widehat{a}_𝐤^{}\widehat{a}_{𝐤+𝐪+𝐪^{}}[\mathrm{exp}\{i(\epsilon _𝐤t\epsilon _{𝐤+𝐪+𝐪^{}}t^{}i\epsilon _{𝐤+𝐪}(tt^{})\}$$ $$\mathrm{exp}\left\{i(\epsilon _𝐤t^{}\epsilon _{𝐤+𝐪+𝐪^{}}t+i\epsilon _{𝐤+𝐪^{}}(tt^{})\}\right].$$ We construct a modified perturbation theory with the help of excluding an integrable part of the interaction. For this purpose we expand the operator $`\widehat{H}_{int}(t)`$ in two parts, $`\widehat{H}_{int}^{(0)}(t)`$ and $`\widehat{H}_{int}^{(1)}(t)`$, where $$\widehat{H}_{int}^{\left(0\right)}\left(t\right)=\widehat{b}^{}\underset{𝐪}{}g_𝐪\widehat{\rho }_𝐪f_𝐪\left(t\right)+\widehat{b}\underset{𝐪}{}g_𝐪^{}\widehat{\rho }_𝐪^{}f_𝐪^{}\left(t\right)$$ $$\widehat{H}_{int}^{\left(1\right)}\left(t\right)=\widehat{H}_{int}\left(t\right)\widehat{H}_{int}^{\left(0\right)}\left(t\right).$$ Here the function $`f_𝐪(t)`$ must be unimodular to preserve the interaction intensity. Obviously, the operators $`\underset{𝐪}{}g_𝐪\widehat{\rho }_𝐪f_𝐪(t)`$ defined at different times, obey the commutation relations. Then, by virtue of the above consideration, the equation $$i\frac{d}{dt}\left|t\right)=\widehat{H}_{int}^{\left(0\right)}\left(t\right)\left|t\right)$$ acquires an exact solution $$\left|t\right)=e^{i\widehat{\chi }\left(t\right)}|h,𝐤_0>$$ (14) where $`\widehat{Q}`$ has the previous form (1) and $$h_𝐪=ig_𝐪\underset{0}{\overset{t}{}}𝑑t^{}f_𝐪\left(t^{}\right)e^{i\omega t^{}}$$ $$\widehat{\chi }\left(t\right)=\frac{i}{2}\underset{0}{\overset{t}{}}\left\{\widehat{\dot{Q}}^{}\left(t^{}\right)\widehat{Q}\left(t^{}\right)\widehat{Q}^{}\left(t^{}\right)\widehat{\dot{Q}}\left(t^{}\right)\right\}𝑑t^{}.$$ The solution (14) can be rewritten as $`|t)=\widehat{U}_0(t)|0,𝐤_0>`$, where we introduce a zero-th order evolution operator $$\widehat{U}_0\left(t\right)=\mathrm{exp}\left\{\widehat{Q}\left(t\right)\widehat{b}^{}\widehat{Q}^{}\left(t\right)\widehat{b}i\widehat{\chi }\left(t\right)\right\}.$$ There are the following useful commutation relations: $$[\widehat{b},\widehat{U}_0\left(t\right)]_{}=\widehat{U}_0\left(t\right)\widehat{Q}\left(t\right)[\widehat{b},\widehat{U}_0^{}\left(t\right)]_{}=\widehat{U}_0^{}\left(t\right)\widehat{Q}\left(t\right)$$ $$[\widehat{b}^{},\widehat{U}_0\left(t\right)]_{}=\widehat{U}_0\left(t\right)\widehat{Q}^{}\left(t\right)[\widehat{b}^{},\widehat{U}_0^{}\left(t\right)]_{}=\widehat{U}_0^{}\left(t\right)\widehat{Q}^{}\left(t\right).$$ Let us introduce a new representation for the vector of state and for operators: $$|t>=\widehat{U}_0^{}\left(t\right)|t)\stackrel{~}{A}=\widehat{U}_0^{}(t\left)\widehat{A}\widehat{U}_0\right(t).$$ The new vector of state obeys the equation $$i\frac{d}{dt}|t>=\stackrel{~}{H}_{int}^{\left(1\right)}\left(t\right)|t>$$ which can be solved with the help of a standard technique using the T-exponent $$|t>=\mathrm{Texp}\{i\underset{0}{\overset{t}{}}dt^{}\stackrel{~}{H}_{int}^{\left(1\right)}\left(t^{}\right)\}|0,𝐤_0).$$ (15) If the choice of the function $`f_𝐪(t)`$ ensures the rapid convergence to the series (15), formula (14) gives a good approximation for the vector of state. In this case we can evaluate a wide set of physical characteristics with sufficient accuracy. As an example, we calculate the density matrix for the particle, for which the exact expression is given by the formula $$\mathrm{\Gamma }(𝐱,𝐱^{},t)=<t\left|\stackrel{~}{\psi }^{}(𝐱,t)\stackrel{~}{\psi }(𝐱^{},t)\right|t>.$$ (16) Here the usual wave operators are introduced, namely, $$\stackrel{~}{\psi }(𝐱,t)=\widehat{U}_0^{}\left(t\right)\widehat{\psi }(𝐱,t)\widehat{U}_0\left(t\right)\widehat{\psi }(𝐱,t)=\underset{𝐤}{}\widehat{a}_𝐤\mathrm{exp}\left\{i\mathrm{𝐤𝐱}i\epsilon _𝐤t\right\}$$ where $`\epsilon _𝐤`$ \- is an energy of the particle posessing momentum $`𝐤`$. The further consideration will be more convenient if the particle- oscillator interaction began at any incident time $`t_0<0`$ when the oscillator was found in the ground state. Let us define the density matrix at $`t=0`$. In the first approximation we can set $`|t>|0,𝐤_0)`$. In this case $$\mathrm{\Gamma }(𝐱,𝐱^{},t)(0,𝐤_0\left|\widehat{U}_0^{}\left(0\right)\widehat{\psi }^{}(𝐱,0)\widehat{\psi }(𝐱^{},0)\widehat{U}_0\left(0\right)\right|0,𝐤_0).$$ (17) Using relation (7) we have $`\widehat{U}_0(0)|0,𝐤_0)=e^{i\widehat{\chi }(0)}|h,𝐤_0>`$, where $`\widehat{Q}`$ is defined as in (1) with $$h_𝐪=h_𝐪\left(0\right)h_𝐪\left(t\right)=ig_𝐪\underset{t_0}{\overset{t}{}}f_𝐪\left(t^{}\right)e^{i\omega t^{}}𝑑t^{}t>t_0.$$ Now we apply relations (12) to obtain the formula similar to (11): $$\widehat{\psi }(𝐱^{},0)\widehat{U}_0\left(t\right)|0,𝐤_0)=\mathrm{exp}\{i𝐤_0𝐱^{}i\mathrm{\Phi }\left(𝐱^{}\right)\}\left|\alpha (𝐱^{},0)\right)\left|vac_p\right)$$ (18) where $$\alpha (𝐱,t)=\underset{𝐪}{}h_𝐪\left(t\right)e^{i\mathrm{𝐪𝐱}}$$ $$\mathrm{\Phi }\left(𝐱\right)=\underset{t_0}{\overset{0}{}}\mathrm{Im}\left[\dot{\alpha }^{}(𝐱,t^{})\alpha (𝐱,t^{})\right]𝑑t^{}.$$ Substituting (18) into (17) we obtain $$\mathrm{\Gamma }(𝐱,𝐱^{},\mathrm{\hspace{0.17em}0})e^{i𝐤_0𝐱+i𝐤_0𝐱^{}}\times $$ $$\mathrm{exp}\{i\mathrm{\Phi }\left(𝐱\right)i\mathrm{\Phi }\left(𝐱^{}\right)\frac{1}{2}[|\alpha (𝐱,0)|^2+|\alpha (𝐱^{},0)|^22\alpha ^{}(𝐱,0)\alpha (𝐱^{},0)]\}.$$ (19) Note, that in the case $`g_𝐪=g\mathrm{\Delta }(𝐪𝐪_0)`$, the phase $`\mathrm{\Phi }(𝐱)=const`$ and formula (19) is simlified. The work was partly supported by the Russian Foundation for Basic Research (grant no 97-02-16058).
warning/0003/astro-ph0003275.html
ar5iv
text
# Old models for Cygnus X-1 and AGN Recently, there appeared many papers devoted to the modeling of X-ray properties of Cygnus X-1 and other black hole accretion disk candidates: e.g. J.Poutanen, J.Krolik, F.Ryde, MNRAS 292 (1997) L21; E.Agol, J.Krolik, ApJ 507 (1998) 304; A.Beloborodov, in “High Energy Processes in Accreting Black Holes”, ASP Conf. Series, 161, (1999), p.295. The goal of this electronic publication is to draw attention to our old papers where many ideas of recent discussions were anticipated (hot coronae, Comptonization, photon damping of waves, particle acceleration and matter ejection from accretion disks with large scale poloidal magnetic fields, etc.) A hot corona around a black-hole accretion disk as a model for Cygnus X-1 G. S. Bisnovatyi-Kogan and S. I. Blinnikov Institute for Space Research, USSR Academy of Sciences, Moscow (Submitted April 15, 1976) Pis’ma Astron. Zh. 2, 489-493 (October 1976) Sov.Astron.Lett. 2, 191-193 (Sep.-Oct. 1976) Heat transfer in the region of maximum energy release of an accretion disk will take place mainly by convection, serving to enhance the turbulence and to generate a powerful acoustic flux. The hard X-rays emitted by Cyg X-1 ($`E<200`$ keV) might result from Comptonization of soft photons in a corona formed around the disk through this heating. PACS numbers: 98.60.Qs, 97.70.Ss With high probability, the X-rays emitted by the source Cygnus X-1 come from a black hole that is in a regime of disk accretion. The theory of disk accretion has been developed by several authors –. But difficulties arise if one seeks to explain the spectrum of Cyg X-1 on the basis of this theory. The source has a luminosity $`L<10^{38}`$ erg/sec $`0.1L_c`$, where $`L_c=1.310^{38}M/M_{}`$ is the critical Eddington luminosity. For this luminosity the theoretical spectrum should fall off sharply in the energy range $`h\nu 7`$ keV, whereas the observed spectrum extends up to energies as high , as $`200`$ keV, or even beyond . To overcome this difficulty, models have been proposed wherein the properties of the radiating regions are essentially determined by the observations. An attempt has been made to construct a self-consistent model for Cyg X-1 that would account for the hard part of the radiation. This model, which neglects radiation pressure from the very outset, appears to have an internal contradiction in that it implies a disk as thick as its radius and drift velocity comparable to the orbital velocity. Hard radiation up to 200-keV energy can be present in the spectrum of an accretion disk only in the event that regions with hot $`(T_e10^9`$ K) electrons participate in forming the spectrum; harder photons would require still more energetic particles. In an analysis of accretion disks we have shown that hot regions or coronae with $`T_e10^9`$ K will form around them. The standard theory presupposes that radiative heat conduction is responsible for energy transfer in the vertical direction. In the region of maximum energy release, $`PP_rP_g`$, $`\kappa =\kappa _{\mathrm{es}}`$, and the vertical structure can easily be determined analytically. The density in this region is found to be independent of $`z`$, while the temperature declines toward higher $`z`$. Such a situation will evidently be convectively unstable, since the entropy will decrease in the vertical direction. The onset of convective instability will serve to equalize the entropy. The mean density in the disk will become approximately an order of magnitude higher than the density of the disk in the standard theory. The convective heat flow $`Q_{\mathrm{conv}}`$, the convection velocity $`v`$, and the excess $`\mathrm{\Delta }T`$ of the temperature gradient above the adiabatic gradient (we shall take the mixing length $`l`$ to be the half-thickness $`z_0`$ of the disk) are given by the equations $$Q_{\mathrm{conv}}=C_p\rho v(z_0/2)\mathrm{\Delta }T,$$ (1) $$v=\left[\frac{Q_{\mathrm{conv}}(1+4P_r/P_g)}{2\rho C_pT}\frac{GM}{R^3}\right]^{1/3}z_0^{2/3},$$ (2) $$\mathrm{\Delta }T=\left(\frac{4Q_{\mathrm{conv}}}{\rho C_p}\right)^{2/3}\left[\frac{T}{(1+4P_r/Pg)GM}\right]^{1/3}Rz_0^{5/3}.$$ (3) Hence we readily find that the excess $`\mathrm{\Delta }T`$ in the temperature gradient amounts to no more than $`20\%`$ of $`T`$, while the convection velocity is $`v=310^8`$ cm/sec for $`r=10r_g`$, a value close to the velocity $`v_s`$ of sound in this region. The heat flow is carried mainly by convection $`(QQ_{\mathrm{conv}})`$; because of the high density the radiative flux is equal to $`20\%`$ of the total flux. In the convective disk the flow $`Q_{\mathrm{ac}}`$ of acoustic energy in the vertical direction is given by $$Q_{\mathrm{ac}}\rho v^3(v/v_s)^510^{2122}\text{erg}\text{sec}^1\text{cm}^2$$ (4) and is a quantity of the same order as the total energy flux. An amount of order $`(P_g/P_r)Q_{\mathrm{ac}}`$ is expended in heating the outer layers to an optical depth $`\tau <1`$. One other mechanism serves to heat these layers . A particle in a region near the surface of the disk that is transparent to radiation will be subject to the influence of radiation from the whole disk, and not only to the local radiation pressure gradient. The force of the radiation pressures will accelerate particles in the transparent region above the disk. Calculations of the equations of motion for particles subject to radiative, centrifugal, and gravitational forces show that in the region with $`\tau <1`$ the particles (protons and electrons) will acquire vertical velocities corresponding to the temperatures $`(18)10^8`$ K of protons for $`L0.1L_c`$. Turbulent relaxation of the particle motions in the corona will tend to equalize the mean energies of the ions and electrons and to form a quasi-Maxwellian particle velocity distribution. Thus the combined action of the two heating mechanisms we have described will produce around the accretion disk a hot corona with $`T_e10^9`$ K for $`L0.1L_c`$. The existence of an analogous corona has been postulated phenomenologically by Price and Liang . We now estimate the density $`\rho _{\mathrm{co}}`$ of the corona and the amount of material it contains. Since the gas pressure varies continuously with transition from the photosphere to the corona, we readily find that at the base of the corona $$\rho _{\mathrm{co}}\rho _sT_s/T_{\mathrm{co}}10^2\rho _s10^5\text{g/cm}^3,$$ (5) where $`\rho _s`$, $`T_s`$ denote the density and temperature in the photosphere of the disk for the parameters of the source Cyg X-1 in the region of maximum energy release. The surface density of the corona is determined by the condition $`\tau _{\mathrm{es}}1`$ and is $`2\text{g/cm}^2`$, which is more than an order of magnitude lower than the density of the opaque disk. The hot corona readily affords an explanation of the peculiarities of the Cyg X-1 spectrum. The soft X-rays at $`h\nu <7`$ keV, which comprise $`70\%`$ of the total flux, are formed in the photosphere of the opaque disk. Some of the radiation $`(10\%)`$ will pass through the hot corona, and Comptonization will generate hard radiation up to $`h\nu 3kT_e200`$ keV, which will amount to $`30\%`$ of the total flux. In Cyg X-1 the luminosity varies around $`L=0.1L_c`$. Under these conditions the region of the disk with $`P_rP_g`$, where convective heating is important, will be spatially small and will convert $`10\%`$ of the soft radiation flux, in accord with the requirements that the observations impose on the model. The changes in the spectrum as the luminosity of Cyg X-1 varies exhibit the following characteristic behavior. As the total energy flux rises, the radiation in the soft part of the spectrum ($`h\nu <7`$ keV) increases, but in the hard range ($`h\nu >10`$ keV) it remains almost constant, or perhaps may even decrease slightly. We shall assume that the variations in the luminosity are associated with fluctuations in the power of accretion. As the mass flow $`\dot{M}`$ rises in the region with $`P_rP_g`$ the fraction of the acoustic flow (4) expended in heating the corona will decrease: $$(P_g/P_r)Q_{\mathrm{ac}}\dot{M}^1.$$ (6) For $`L0.1L_c`$, when acoustic heating predominates, the rise in $`\dot{M}`$ may cause some decrease in the heating of the corona and in the amount of hard radiation. At the same time the flux in the soft range is determined by the radiation of the disk photosphere and will increase with $`\dot{M}`$. It is worth noting that in strong bursts of luminosity, when $`L`$ reaches about $`0.3L_c`$, the heating will begin to be governed by radiation-pressure forces, so the temperature of the corona and thereby also the power of the hard radiation should increase along with the rise in $`\dot{M}`$ and in the total energy flux. The observational evidence for the spectrum at $`h\nu >150`$ keV is considerably less reliable than for the soft range. For instance, Baker et al. have measured the radiation to $`h\nu 10`$ MeV, but find an observed signal of only $`1\%`$ of the background (they also report a deficiency of $`\gamma `$-ray photons in the direction of Cyg X-1). In the range up to 600 keV, Haymes and Harnden find a break in the spectrum at $`h\nu 150`$ keV, and the spectral index $`\alpha `$ changes from $`\alpha =1.9`$ at $`h\nu <124`$ keV to $`\alpha =3.1`$ at $`h\nu >154`$ keV. Even if a corona with $`T10^9`$ K is present in the disk accretion model, radiation cannot be formed at $`h\nu =2002000`$ keV. Electrons with $`T_e=10^{10}`$ K or fast nonthermal electrons would be needed for that purpose. There are in fact two ways to form such fast electrons. Both would require the presence of a magnetic field in the disk. A magnetic field could exist in the disk either through twisting of the lines of force by differential rotation , or through infall onto the black hole of magnetized material having a small angular momentum . In the latter case a poloidal magnetic field would be generated. In the binary system containing the source Cyg X-1, some of the material flowing from the giant star is dispersed in space near the system. The attraction of the black hole will not only produce an accretion disk. A small proportion of material having a low angular momentum will fall into the black hole and be decelerated in the disk. If this deceleration takes place at radii of $`(1030)r_g`$ and if a thin collisionless shock wave is formed (as is very likely in the presence of an azimuthal magnetic field) wherein the kinetic energy is transformed into thermal energy and $`T_eT_i`$, then hot electrons with $`T=10^{10}10^{12}`$ K will appear. The inverse Compton mechanism of interaction of the disk radiation with these electrons can lead to the generation of hard radiation with $`h\nu 2002000`$ keV or even higher energies. Another mechanism for producing fast particles is analogous to the pulsar process. If magnetized matter with low angular momentum falls into the black hole (in addition to the disk accretion), a strong poloidal magnetic field will arise . By analogy to pulsars , rotation will generate an electric field of strength $`E(v/c)B`$ in which electrons are accelerated to energies $`\epsilon R(v/c)Be310^4[B/(10^7\text{Gauss})]`$ Mev where $`v/c0.1`$ and $`R10^7`$ cm is the characteristic scale. In a field $`B10^7`$ Gauss, such electrons will generate synchrotron radiation with energies up to $`10^5`$ keV. Just as in pulsars, it would be possible here for $`e^+e^{}`$ pairs to be formed and to participate in the synchrotron radiation. The authors express their appreciation to Ya.B. Zel’dovich, A. F. Illarionov, and I. S. Shklovskii for helpful discussions and valuable comments. 1. N. I. Shakura, “A disk model for accretion of gas by a relativistic star in a close binary system”, Astron. Zh. 49, 921-929 (1972) \[Sov. Astron.16,756-762 (1973)\]. 2. J. E. Pringle and M.J.Rees, “Accretion disk models for compact X-ray sources”, Astron. Astrophys. 21, 1-9 (1972). 3. N. I. Shakura and R. A. Syunyaev, “Black holes in binary systems: observational appearance”, Astron. Astrophys. 24, 337-355 (1973). 4. I. D. Novikov and K. S. Thorne, “Black-hole astrophysics”, in: Black Holes (Les Houches lectures, Aug. 1972), Gordon and Breach (1973), pp.343-450. 5. F. Frontera and F. Fuligni, “Energy-spectrum variability of Cyg X-1 in hard X-rays”, Astrophys. J. 196, 597-599 (1975). 6. G. F. Carpenter, M. J. Coe, A. R. Engel, and J. J. Quenby, “Ariel 5 hard X-ray measurements of galactic and extragalactic source spectra”, Proc.14th Intl. Cosmic Ray Conf. (Munich) 1, 174-179 (1975). 7. R. C. Haymes and F. L. Harnden, “Low-energy $`y`$ radiation from Cygnus” Astrophys. J. 159, 1111-1114 (1970). 8. K. E. Baker, R. L. Lovett, K. J. Orford, and D. Ramsden, “Gamma rays of 1-10 MeV from the Crab and Cygnus regions”, Nature Phys. Sci. 245, 18-19 (1973). 9. K. S. Thorne and R. H. Price, “Cyg X-1: an interpretation of the spectrum and its variability”, Astrophys. J. 195, L101-L105 (1975). 10. R. H. Price and E. P. Liang, Preprint (1975). \[published in Astrophys. J., 218, Nov. 15, 1977, 247-252.\] 11. S. L. Shapiro, A. P. Lightman, and D. M. Eardley, “A two-temperature accretion-disk model for Cyg X-l”, Astrophys. J. 204, 187-199 (1976). 12. G. S. Bisnovatyi-Kogan and S. I. Blinnikov, Preprint Inst. Kosmich. Issled. Akad. Nauk SSSR No. 271 (1976); Astron. Astrophys. (in press). \[published in A&A 1977, 59, 111-125\] 13. M. Schwarzschild, Structure and Evolution of the Stars, Princeton Univ. Press. (1958). 14. L. Biermann and R. Lust. “Nonthermal phenomena in stellar atmospheres”, in: Stellar Atmospheres, Univ. Chicago Press (1960), Chap. 6. 15. D. R. Parsignault, A. Epstein, J. E. Grindlay, E. J. Schreier, H. Schnopper, H. Gursky, Y. Tanaka, A. C. Brinkman, J. Heise, 1. Schrijver, R. Mewe, E. Gronenschild, and A. den Boggende, “ANS observations of Cyg X-1”, Astrophys. Space Sci. 42, 175-184 (1976). 16. S. S. Holt, E. A. Boldt, L. J. Kaluzienski, and P. J. Serlemitsos, “Observations of a new transition in the emission from Cyg X-1”, Nature 256, 108-109 (1975). 17. L. A. Pustil’nik and V. F. Shvartsman, “Possible influence of magnetic fields on the structure of a plasma accretion disk in binary systems”, in: Gravitational Radiation and Gravitational Collapse (IAU Sympos. No. 64), Reidel (1974), p. 213. 18. D. M. Eardley and A. P. Lightman, “Magnetic viscosity in relativistic accretion disks”, Astrophys. J. 200, 187-203 (1975). 19. G. S. Bisnovatyi-Kogan and A. A. Ruzmaikin, “The accretion of matter by a collapsing star in the presence of a magnetic field”, Astrophys. Space Sci. 42, 401-424 (1976). 20. P. Goldreich and W. H. Julian, “Pulsar electrodynamics”, Astrophys. J., 157, 869-880 (1969). Models for the X-ray brightness fluctuations in Cygnus X-1 and active galaxy nuclei G. S. Bisnovatyi-Kogan and S. I. Blinnikov Institute for Space Research, USSR Academy of Sciences, Moscow (Submitted April 7, 1978) Pis’ma Astron. Zh. 4, 540-543 (December 1978) Sov.Astron.Lett. 4, 290-291 (Nov.-Dec. 1978) The X-ray brightness fluctuations induced in Cyg X-l and active galaxy nuclei by convection and turbulence in the sub-photospheric layers are discussed in terms of the disk-accretion and supermassive-star models. The variability time scales should be comparable with those observed, but in the case of supermassive stars and disks it is difficult to obtain the observed amplitude of the fluctuations. PACS numbers: 98.70.Qy, 98.50.Rn, 97.10.Cv 1. One noteworthy property of the X-ray source Cygnus X-l is the variability of its flux on time scales ranging from milliseconds to a fraction of a second . This variability is generally attributed to the presence of a turbulent accretion disk in the Cyg X-l system. In particular, the flux variations have been explained by the rotation of a hot spot or alternatively, by the onset of instability in a zone where radiation pressure predominates . In this letter we shall consider another model for the variability. We shall outline more fully what properties of the flux variations are to be expected on the accretion disk model as a direct result of the presence of turbulence, and how convective instability should develop in a region of high radiation pressure . Acoustic waves generated in the convection zone will escape into optically thin layers, and will not only induce variable soft X-rays in the photosphere, but will also be responsible for variable heating of the corona. Comptonization of the photospheric radiation by the hot electrons under conditions where both temperature and density are variable will lead to flux variations in the hard range, $`h\nu >`$ 5 keV. This X-ray fluctuation mechanism, involving the emergence of waves into transparent layers, evidently is of the same wave nature as the variability of the ultraviolet excess in stars experiencing intensive convection, such as T Tauri and UV Ceti. 2. The waves escaping into the transparent layers and producing variable radiation in the photosphere and corona will occupy a rather narrow frequency band. Physically, the reason for this circumstance is that media with a high radiation pressure and a nonuniform distribution of plasma along the $`z`$ coordinate (across the disk) will serve as efficient filters, isolating a characteristic frequency range from the broader spectrum generated by convection and turbulence. Waves of low frequency and a wavelength exceeding the scale height of the atmosphere will not escape outside but will induce oscillations of the coronal atmosphere as a whole. On the other hand, under conditions where radiation pressure predominates, high-frequency waves will experience very severe damping because of radiative friction, and their role in heating the corona will be insignificant. High frequencies may, however, appear in the observations either through nonlinear transformation of low-frequency waves or due to inadequate processing of the observational data . We have examined elsewhere the propagation of waves through a medium with strong radiation pressure, followed by their escape into the atmosphere. Waves emerging into the transparent layers will have a phase and group velocity equal to the velocity of sound in gas, $`v_g=(\gamma P_g/\rho )^{1/2}=(\gamma T)^{1/2}`$, where $`\gamma =`$5/3 or $`\gamma =`$ 1 according as scattering or absorption predominates, $``$ is the gas constant, and the temperature $`T`$ of the equilibrium atmosphere is approximately equal to the temperature $`T_\mathrm{e}`$ of the photosphere. The characteristic frequency $`\omega _\mathrm{c}`$ of waves emerging into the atmosphere is given by the expression $$\omega _\mathrm{c}=\left(\frac{\gamma }{T}\right)^{1/2}g\left(1\frac{H}{H_c}\right)\mathrm{sec}^1.$$ (7) For the accretion disk model, the gravitational acceleration $`g`$ at radius $`r`$ is equal to $$g=\frac{GM}{r^2}\frac{h_0}{r},$$ (8) where the characteristic thickness is $$h_0=3\frac{L}{L_c}R_0\left[1\left(\frac{R_0}{r}\right)^{1/2}\right],$$ (9) and $`R_0`$ is the radius of the inner edge of the disk. The quantity $`H/H_\mathrm{c}=\kappa _0H/gc`$ represents the ratio of the radiation pressure force to the gravity; $`\kappa _0`$ is the opacity, including both absorption and scattering . In a spherically symmetric star of luminosity $`L`$ and radius $`R`$ we will have $`g=GM/R^2`$, $`H=L/4\pi R^2`$, and $`H/H_c=L/L_c`$, where $`L_c=4\pi cGM/\kappa _0`$ is the Eddington limiting luminosity. 3. To calculate the characteristic frequencies of the fluctuations for Cyg X-l, we shall adopt the convective accretion-disk models given in a previous paper and by Shakura et al. . For a black hole of mass $`M=10M_{}`$, we find that in both models as the luminosity rises from $`0.1L_c`$ to $`0.3L_c`$ there is little change in the characteristic frequency $`\omega _c`$ corresponding to the equilibrium temperature, as given by Eq. (1); the values of $`\omega _\mathrm{c}`$ are confined to the range 5-40 msec. Note that if a corona with $`T_c10^2T_e`$ is present, waves whose frequency is $`\omega =\omega _\mathrm{c}`$ or even somewhat lower will be able to escape. The frequency range mentioned is in good accord with the observed time-scales of variability. Our mechanism can yield fluctuations weakly correlated in time, and can simulate the white noise derived from observational analysis of the brightness fluctuations of Cyg X-l. Despite the good agreement between the theoretical time characteristics and the observations, analysis of small oscillations in our model still does not suffice to explain the amplitudes of the variability we observe, because pulsations should, in general, occur independently in regions whose size is comparable with the velocity of sound multiplied by the pulsation period, or about one-tenth the diameter of the zone of maximum energy production. Small pulsations should accordingly be smoothed out. However, the mechanism we are proposing could operate in a highly nonlinear regime. Strong nonlinearity of the waves would be expected in the light of theoretical estimates , and is essential to the very existence of the corona. The frequencies obtained from linear analysis correspond to the characteristic growth times of strong nonlinear flares. This problem awaits further theoretical treatment. Indeed, we would point out that not even the observational situation is fully clear . An evaluation of the proportion $`\eta `$ of the acoustic flux that emerges into the corona and produces heating shows that as the luminosity rises by a factor of 3 from 0.1 $`L_c`$ to 0.3 $`L_c`$, the value of $`\eta `$ will drop by a factor of 5-10. Thus, in accord with our previous suggestion , as the luminosity increases the power of the corona will remain approximately the same, or may even decline somewhat. This situation will prevent the hard X-ray flux ($`h\nu >5`$ keV) of Cyg X-l from changing appreciably as the total luminosity varies, in agreement with the observations . 4. Multicolor photometry of the nuclei of certain galaxies has revealed the presence of comparatively short-period brightness fluctuations in the nucleus of the Seyfert galaxy NGC 4151 ($``$ 130 days) and in the BL Lacertae object OJ 287 ($``$ 180 days). There is presently no consensus as to the nature of active galaxy and quasar nuclei. Three models have been explored : a) a dense star cluster; b) a supermassive star; c) disk accretion onto a massive black hole. The variability mechanism proposed here can operate in the two models b and c, for in both cases the subphotospheric layers will be strongly convective, resulting in the formation of a corona with variable heating. Using Eq. (1) to estimate the characteristic period of the fluctuations along with the standard model for a supermassive star we find that for a mass $`M=10^8M_{}`$ and a radius $`100R_\mathrm{g}`$ ($`R_g=2GM/c^2`$ is the gravitational radius), the period corresponding to $`\omega _c`$ would be $`160`$ days, a value consistent with observation. On the accretion disk model, such a period would prevail in the zone of maximum energy production for a black hole of $`M=10^8M_{}`$, if the luminosity $`L=0.1L_c`$. In galaxy nuclei the fast component typically exhibits an approximately constant period in conjunction with sharp changes in phase , a behavior which, it would seem, accords with a convective wave origin for such fluctuations. It is worth emphasizing, however, that both in the supermassive star model and in the model of an accretion disk around a supermassive black hole, the ratio of the radiation pressure to the gas pressure is far higher than in the case of accretion onto a black hole of stellar mass. Acoustic waves will therefore be damped much more strongly. Our calculations indicate that in this event the emergent acoustic flux will comprise no more than 1$`\%`$ of the flux generated at large optical depth - well below the variability amplitude observed. The supermassive star model would be supported if strictly periodic brightness oscillations of constant phase were detected, associated with the rotation or with pulsations of such a star as a whole, as might be the case for NGC 4151. But analysis of observations of several variable nuclei does not reveal any strict periodicities . Most likely the observed variability should be modeled by a random process (see, however, Ozernoi et al. ). The model of a dense star cluster with frequent supernova outbursts offers the best fit, in our opinion, to the irregular variability of galaxy nuclei. 1. E. Boldt, “X-ray signatures: new time scales and spectral features,” in: Eighth Texas Sympos. on Relativistic Astrophysics. Ann. New York Acad. Sci. 302, 329-348 (1977). 2. R. A. Syunyaev, “Variability of X-rays from black holes with accretion disks,” Astron.Zh. 49, 1153-57 (1972) \[Sov. Astron. 16, 941-944 (1973)\]. 3. A. P. Lightman and D. M. Eardley, “Black holes in binary systems: instability of disk accretion,” Astrophys. J. 187, L1-L3 (1974). 4. N. Shibazaki and R. Hoshi. “Structure and stability of an accretion disk around a black hole,” Progr. Theor. Phys. 54, 706-718 (1975). 5. N. I. Shakura and R. A. Syunyaev, “A theory of the instability of disk accretion onto black holes,” Mon. Not. R. Astron. Soc. 175, 613-632 (1976). 6. G. S. Bisnovatyi-Kogan and S. I. Blinnikov, “A hot corona around a black-hole accretion disk as a model for Cyg X-l,” Pis’ma Astron. Zh. 2, 489-493 (1976) \[Sov. Astron Lett. 2, 191-193 (1977)\]. 7. G. S. Bisnovatyi-Kogan and S. I. Blinnikov, “Disk accretion onto a black hole at subcritical luminosity,” Astron. Astrophys. 59, 111-125 (1977). 8. M. C. Weisskopf and P. G. Sutherland, “On the physical reality of the millisecond bursts in Cyg X-l,” Astrophys. J. 221, 228-233 (1978). 9. G. S. Bisnovatyi-Kogan and S. I. Blinnikov, “Wave propagation in a medium with high radiation pressure” \[in Russian\]. Preprint Inst. Kosmich. Issled. Akad. Nauk SSSR No. 421 (1978). \[Published in Astrophysics 14, 316-325 (1978); 15, 99-107 (1979).\] 10. N. I. Shakura and R. A. Syunyaev. “Black holes in binary systems: observational appearance,” Astron. Astrophys. 24, 337-355 (1973). 11. N. I. Shakura, R. A. Syunyaev, and S. S. Zilitinkevicn, “Turbulent energy transport in accretion disks,” Astron. Astrophys. 62, 179-187 (1978). 12. J. F. Dolan, C. J. Crannell, B. R. Dennis, K. J. Frost, and L. E. Orwig, “Intensity transitions in Cyg X-l observed at high energies from OSO-8,” Nature 267, 813-815 (1977). 13. V. M. Lyutyi and A. M. Cherepashchuk, “H$`\alpha `$ emission variability in the Seyfert nuclei NGC 4151, 3516, 1068” \[in Russian\], Astron. Tsirk. No. 831. 1-3 (1974). 14. E. T. Belokon’, M. K. Babadzhanyants, and V. M. Lyutyi, “Periodicity of the Seyfert galaxy NGC 4151 in the optical,” Astron. Astrophys. Suppl. 31, 383 (1978). l5. V. M. Lyutyi, “Optical variability of OJ 287” \[in Russian\]. Peremennye Zvezdy 20, 243-249 (1976). 16. V. L. Ginzburg and L. M. Ozernoi, “The nature of quasars and active galaxy nuclei,” Astrophys. Space Sci. 50, 23-41 (1977). 17. M. J. Rees, “Quasar theories,” in: Eighth Texas Sympos. on Relativistic Astrophysics. Ann. New York Acad. Sci. 302, 613-636 (1977). 18. Ya. B. Zel’dovich and I. D. Novikov, Relativistic Astrophysics, Vol. 1, Stars and Relativity, Univ. Chicago Press (1971). 19. M. M. Basko and V. M. Lyutyi, “Search for periodic optical variations of the Seyfert nuclei NGC 1275 and 3516,” Pis’ma Astron.Zh. 3, 104-103 (1977) \[Sov. Astron. Lett. 3, 54-56 (1977)\]. 20. W. H. Press, “Flicker noises in astronomy and elsewhere,” Comments Astrophys. Space Phys. 7, 103-119 (1978). 21. L. M. Ozernoi, V. E. Chertoprud, and L. I. Gudzenko, “Comments on the light curve of the quasar 3C 273,” Astrophys. J. 216, 237-243 (1977).
warning/0003/hep-ph0003286.html
ar5iv
text
# Supersymmetric Extension of the Standard Model with Naturally Stable Proton ## I Introduction The standard model (SM) well describes particle physics below the electroweak energy scale. However, various theoretical considerations suggest that some extension of the SM be necessary for physics above that energy scale. Various models therefore have been proposed, some of which being studied extensively. Among them extensions with supersymmetry are considered most plausible. In particular, the minimal supersymmetric standard model (MSSM) is usually treated as their standard theory around the electroweak energy scale. The MSSM inherits most of the successful features of the SM, while the extension being minimal. However, this model suffers one serious setback, which has been often passed over. In the SM, the proton is protected from decay naturally by gauge symmetry. On the other hand, in the MSSM, the gauge symmetry allow the interactions of dimension four which do not conserve baryon and/or lepton numbers. Unless there exists some reason to forbid these interactions, the proton decays in an unacceptably short time. Therefore, a discrete symmetry is usually imposed on the MSSM through $`R`$ parity, which is however merely an ad hoc symmetry. A convincing reason for the proton stability could be provided by an extra gauge symmetry. Although such a symmetry around the electroweak energy scale is subjected to many phenomenological constraints, they still show room to allow a U(1) gauge symmetry. Several supersymmetric models with an extra U(1) symmetry therefore have been discussed , aiming at natural explanation for a long lifetime of the proton. However, these models are accompanied by some arbitrariness in construction, which might reduce reliability of the reasoning for the proton stability. In this paper, the supersymmetric extension of the SM with an extra U(1) gauge symmetry is studied within the framework of a model coupled to $`N=1`$ supergravity. In addition to the proton lifetime, the MSSM involves problems on the neutrino masses and the linear coupling of Higgs superfields noted later. Requiring a model to solve these problems consistently in a minimal extension, its particle contents and superpotential are determined rather uniquely . In sizable ranges of the model parameters, the scalar potential appropriately gives a vacuum of SU(3)$`\times `$U<sub>EM</sub>(1) gauge symmetry below the electroweak energy scale. Phenomenological predictions of the model are compatible with experimental results. A typical mass scale of scalar particles is of order 1 TeV, which can account for the smallness of the electric dipole moments (EDMs) of the neutron and the electron, another problem of the MSSM. The energy dependencies of the model parameters are also discussed by analyzing renormalization group equations (RGEs). Taking the masses-squared of scalar fields all positive at a high energy scale, those for some Higgs fields become sufficiently small at lower energy scales to induce the breakdowns of the extra U(1) and electroweak gauge symmetries. In the model coupled to $`N=1`$ supergravity the masses-squared and the trilinear coupling constants for scalar fields are considered to respectively have universal values at a very high energy scale. This scenario can be realized in this model. In constructing the model, we take into account the problems of the neutrino masses and the Higgs linear coupling as well as that of the proton stability. The former is raised by non-vanishing masses of the neutrinos suggested from experiments for atmospheric and solar neutrinos, such as at the Super-Kamiokande . The MSSM or the SM can have Yukawa couplings for neutrino Dirac masses, if right-handed neutrinos are naively included. However, these fields are inert for the transformations of the gauge groups and their existence is not prescribed by the model. Furthermore, the extreme lightness of the neutrinos may require some explanation. Although this lightness could be attributed to large Majorana masses of the right-handed neutrinos, their origin is not clarified. The latter problem is posed by a mass parameter of the $`\mu `$ term, a linear coupling of the Higgs superfields in the superpotential of the MSSM , which is indispensable for correct breaking of electroweak gauge symmetry. This mass parameter $`\mu `$ should have a magnitude of order the electroweak energy scale. The other mass parameters in the model are traced back to supersymmetry-soft-breaking terms of the Lagrangian and thus related to the gravitino mass which may be of order the electroweak energy scale. On the other hand, the $`\mu `$ parameter is contained in the supersymmetric term and its magnitude may be given arbitrarily. It is natural that there is no mass parameter in the superpotential and the role of the $`\mu `$ term is assumed by another effective $`\mu `$ term. These problems could also be solved by introducing an extra U(1) gauge symmetry. Imposing a new gauge symmetry yields chiral and trace anomalies within the particle contents of the MSSM. For canceling the anomalies, new superfields are necessarily incorporated, among which those for right-handed neutrinos and an SU(2)-singlet Higgs boson may be included. The extra U(1) symmetry could be broken above the electroweak energy scale by a vacuum expectation value (v.e.v.) of this Higgs boson, which may provide a source of Majorana masses for the right-handed neutrinos and the effective $`\mu `$ parameter. This paper is organized as follows. In Sect. 2 we construct a model, in which the proton is adequately stable, the ordinary neutrinos have non-vanishing but small masses, and the effective $`\mu `$ term is contained. In Sect. 3 the vacuum structure of the model is discussed, paying particular attention to experimental constraints on an extra neutral gauge boson. In Sect. 4 the behavior of the model parameters for different energy scales is analyzed through the RGEs to examine the radiative breaking of gauge symmetry and the supersymmetry breaking by $`N=1`$ supergravity. Conclusions and discussions are given in Sect. 5. ## II Model Particle contents of the model are constrained by the requirements of proton stability, neutrino masses, and an effective $`\mu `$ term. We also keep the extension of the SM as minimal as possible. The neutrino masses necessitate superfields for right-handed neutrinos, which are denoted by $`N^c`$. For the effective $`\mu `$ term an SU(3)$`\times `$SU(2)$`\times `$U(1) singlet superfield $`S`$ is included. In addition, new colored superfields $`K`$ and $`K^c`$ are necessary for canceling a chiral anomaly, as shown later. In Table I we list the left-handed chiral superfields contained in the model with their quantum numbers under SU(3), SU(2), U(1), and $`\mathrm{U}^{}(1)`$ gauge transformations. The extra U(1) gauge symmetry is denoted by $`\mathrm{U}^{}(1)`$, for which the charges of superfields are expressed as $`Q_Q`$, $`Q_{U^c}`$, etc.. In order not to yield chiral and trace anomalies within the standard gauge symmetries, we assign opposite U(1) charges $`Y_K`$ and $`Y_K`$ to $`K`$ and $`K^c`$. The index $`i`$ $`(i=1,2,3)`$ stands for the generation, while the indices $`j`$ of $`H_1`$ and $`H_2`$, $`k`$ of $`S`$, and $`l`$ of $`K`$ and $`K^c`$ are for possible multiplication to be determined by cancellation of the anomalies. The superpotential should contain the couplings $`H_1QD^c`$, $`H_2QU^c`$, $`H_1LE^c`$, and $`H_2LN^c`$ to generate masses for quarks and leptons. The $`\mu `$ term can be replaced by the coupling $`SH_1H_2`$, provided that the scalar component of $`S`$ has a non-vanishing v.e.v.. The Dirac masses of the neutrinos may be comparable to those of the charged leptons, unless the Yukawa coupling constants are extremely small. However, the ordinary neutrino masses are suppressed by giving large Majorana masses to the right-handed neutrinos, which can be accomplished, without another new field, by including the coupling $`SN^cN^c`$. These couplings provide constraints on the $`\mathrm{U}^{}(1)`$ charges of the superfields: $`Q_{H_1}+Q_Q+Q_{D^c}`$ $`=`$ $`0,`$ (1) $`Q_{H_2}+Q_Q+Q_{U^c}`$ $`=`$ $`0,`$ (2) $`Q_{H_1}+Q_L+Q_{E^c}`$ $`=`$ $`0,`$ (3) $`Q_{H_2}+Q_L+Q_{N^c}`$ $`=`$ $`0,`$ (4) $`Q_S+Q_{H_1}+Q_{H_2}`$ $`=`$ $`0,`$ (5) $`Q_S+Q_{N^c}+Q_{N^c}`$ $`=`$ $`0.`$ (6) If colored superfields are only those which correspond to the quarks of the SM, the $`[\mathrm{SU}(3)]^2\mathrm{U}^{}(1)`$ anomaly-free condition with Eqs. (1) and (2) gives the relation $`Q_{H_1}+Q_{H_2}=0`$. The linear coupling $`H_1H_2`$ are not forbidden in the superpotential, and thus the model inevitably has a mass parameter of unknown origin. Therefore, new colored superfields should be included to solve the problem of the $`\mu `$ term. Although there are various candidates for such superfields, according to the ’minimal’ postulate, we incorporate a pair of superfields in the fundamental representations of the SU(3) group, $`K`$ and $`K^c`$. Then, their fermion components should have large masses from a phenomenological viewpoint, which is fulfilled by allowing the coupling $`SKK^c`$. This coupling leads to another constraint $$Q_S+Q_K+Q_{K^c}=0.$$ (7) The $`[\mathrm{SU}(3)]^2\mathrm{U}^{}(1)`$ anomaly-free condition and Eqs. (1), (2), (5), and (7) fix the number $`n_l`$ of pairs for $`K`$ and $`K^c`$ at three, which agrees with the number of the generation for quarks and leptons. The number $`n_j`$ of pairs for $`H_1`$ and $`H_2`$ and the number $`n_k`$ for $`S`$ are determined by the freedom from chiral anomalies for $`[\mathrm{SU}(2)]^2\mathrm{U}^{}(1)`$, $`[\mathrm{U}(1)]^2\mathrm{U}^{}(1)`$ and a trace anomaly for U(1) with Eqs. (1)-(5), (7). These constraints are satisfied by either of the three sets of numbers and U(1) charge: $`(\mathrm{A})`$ $`n_j=4,n_k=5,Y_K=0,`$ (8) $`(\mathrm{B})`$ $`n_j=3,n_k=3,Y_K=\pm {\displaystyle \frac{1}{3}},`$ (9) $`(\mathrm{C})`$ $`n_j=2,n_k=1,Y_K=\pm {\displaystyle \frac{\sqrt{2}}{3}}.`$ (10) However, the solution (A) does not satisfy the $`\mathrm{U}(1)[\mathrm{U}^{}(1)]^2`$ anomaly-free condition with Eq. (6). The solution (C) gives irrational U(1) charges for $`K`$ and $`K^c`$. The solution (B) is free from the $`\mathrm{U}(1)[\mathrm{U}^{}(1)]^2`$ anomaly, and also satisfies the remaining $`[\mathrm{U}^{}(1)]^3`$ anomaly-free condition. Therefore, a plausible solution is uniquely given by the set (B). The numbers $`n_j`$ and $`n_k`$ again become equal to the number of the generation. A further constraint comes from the stable proton. The allowed value of $`Y_K`$ for the U(1) charges of $`K`$ and $`K^c`$ is now either 1/3 or $`1/3`$. However, the proton stability by gauge symmetry is only achieved for $`Y_K=1/3`$ . For $`Y_K=1/3`$, the particle contents of one generation can be embedded in the fundamental representation of the E<sub>6</sub> group. Unless a discrete symmetry is imposed, the baryon and/or lepton numbers are violated by couplings of dimension four, such as $`U^cD^cK^c`$ and $`LQK^c`$, which induce an unacceptably fast decay of the proton. On the other hand, for $`Y_K=1/3`$, allowed couplings of dimension four are only those which have already been taken into account, i.e. $`H_1QD^c`$, $`H_2QU^c`$, $`H_1LE^c`$, $`H_2LN^c`$, $`SH_1H_2`$, $`SN^cN^c`$, and $`SKK^c`$. Baryon number is conserved while lepton number is not, which is sufficient for the proton stability. The lowest dimension couplings of baryon-number violation are given by the D terms of $`QQU^cE^c`$, $`QQD^cN^c`$, and $`QU^cD^cL`$, which are of dimension six. Under all the anomaly-free conditions and Eqs. (1)-(7), the $`\mathrm{U}^{}(1)`$ charges of the superfields are expressed in terms of two independent variables. All the superfields are triplicated, and the anomalies are canceled in each generation. The generators $`Y^{}`$ and $`Y`$ of $`\mathrm{U}^{}(1)`$ and U(1), respectively, are required to be orthogonal, $`\mathrm{Tr}[Y^{}Y]=0`$. Then, the $`\mathrm{U}^{}(1)`$ charges of the superfields are determined up to a normalization factor. For definiteness, hereafter, the $`\mathrm{U}^{}(1)`$ charges are normalized to the U(1) charges as $`\mathrm{Tr}[Y^2]=\mathrm{Tr}[Y^2]`$, which are shown in Table II. The superpotential which contains all the couplings consistent with gauge symmetry and renormalizability is given by $`W`$ $`=`$ $`\eta _d^{ijk}H_1^iQ^jD^{ck}+\eta _u^{ijk}H_2^iQ^jU^{ck}+\eta _e^{ijk}H_1^iL^jE^{ck}+\eta _\nu ^{ijk}H_2^iL^jN^{ck}`$ (12) $`+\lambda _N^{ijk}S^iN^{cj}N^{ck}+\lambda _H^{ijk}S^iH_1^jH_2^k+\lambda _K^{ijk}S^iK^jK^{ck},`$ where $`\eta _d`$, $`\eta _u`$, $`\eta _e`$, $`\eta _\nu `$, $`\lambda _N`$, $`\lambda _H`$, and $`\lambda _K`$ represent dimensionless constants. Contraction of group indices is understood. In the MSSM without the discrete symmetry through $`R`$ parity, the couplings $`D^cD^cU^c`$, $`LQD^c`$, $`LLE^c`$, $`H_1H_1E^c`$, and $`LH_2`$ are allowed, leading to non-conservation of baryon and lepton numbers. Here, these couplings are forbidden by the U(1) gauge symmetry. The proton decay could only occur through the operators of dimension six, being suppressed at least by a huge mass to the second power. As long as this mass scale is not much smaller than the Planck mass, the proton becomes adequately stable. The couplings of the superpotential are all cubic, and there is no mass parameter. We assume that supersymmetry is broken through the ordinary mechanism based on $`N=1`$ supergravity. Supergravity is spontaneously broken in a hidden sector at the Planck mass scale, and then supersymmetry in an observable sector is broken softly. At lower energy scales, the Lagrangian of the observable sector consists of a supersymmetric part and a supersymmetry-soft-breaking part prescribed by gauge symmetry and superpotential. The soft-breaking part contains mass terms for gauge fermions, and trilinear couplings and mass terms for scalar bosons, $`_𝒮`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\stackrel{~}{m}_3\overline{\lambda }_3\lambda _3+\stackrel{~}{m}_2\overline{\lambda }_2\lambda _2+\stackrel{~}{m}_1\overline{\lambda }_1\lambda _1+\stackrel{~}{m}_1^{}\overline{\lambda }_1^{}\lambda _1^{}\right)`$ (17) $`m_{3/2}(A_d^{ijk}\eta _d^{ijk}H_1^iQ^jD^{ck}+A_u^{ijk}\eta _u^{ijk}H_2^iQ^jU^{ck}+A_e^{ijk}\eta _e^{ijk}H_1^iL^jE^{ck}+A_\nu ^{ijk}\eta _\nu ^{ijk}H_2^iL^jN^{ck}`$ $`+B_N^{ijk}\lambda _N^{ijk}S^iN^{cj}N^{ck}+B_H^{ijk}\lambda _H^{ijk}S^iH_1^jH_2^k+B_K^{ijk}\lambda _K^{ijk}S^iK^jK^{ck})+\mathrm{H}.\mathrm{c}.`$ $`M_{Q^i}^2|Q^i|^2M_{U^{ci}}^2|U^{ci}|^2M_{D^{ci}}^2|D^{ci}|^2M_{L^i}^2|L^i|^2M_{N^{ci}}^2|N^{ci}|^2M_{E^{ci}}^2|E^{ci}|^2`$ $`M_{H_1^i}^2|H_1^i|^2M_{H_2^i}^2|H_2^i|^2M_{S^i}^2|S^i|^2M_{K^i}^2|K^i|^2M_{K^{ci}}^2|K^{ci}|^2.`$ Here $`\lambda _3`$, $`\lambda _2`$, $`\lambda _1`$, and $`\lambda _1^{}`$ represent gauge fermions for SU(3), SU(2), U(1), and U(1), respectively. Scalar bosons are denoted by the same symbols as the corresponding superfields. With $`m_{3/2}`$ being the gravitino mass, the coefficients $`A_d`$, $`A_u`$, $`A_e`$, $`A_\nu `$, $`B_N`$, $`B_H`$, and $`B_K`$ are dimensionless. At high energy scales not much lower than the Planck mass, the masses-squared of scalar fields are all around $`m_{3/2}^2`$ and positive. The trilinear coupling constants for scalar fields are also approximately the same. Around the electroweak energy scale, some of these parameters differ significantly from the high-energy values through large quantum corrections. ## III Vacuum Structure The Lagrangian of our model has SU(3)$`\times `$SU(2)$`\times `$U(1)$`\times `$U(1) gauge symmetry, which must be spontaneously broken down to SU(3)$`\times `$U<sub>EM</sub>(1) symmetry. This breaking could be achieved by v.e.v.s for the scalar components of $`H_1^i`$, $`H_2^i`$, and $`S^i`$. We discuss the vacuum structure of the model by examining the scalar potential. Hereafter, we adopt the same notation for the superfields and their scalar components. Although the scalar potential could contain all of $`H_1^i`$, $`H_2^i`$, and $`S^i`$ and thus its general analysis is complicated, it may be simplified under certain assumptions. If the couplings between different generations are not significant, $`H_2^3Q^3U^{c3}`$ of the third generation has a large coefficient related to the mass of the top quark. The mass-squared of $`H_2^3`$ then receives large negative contributions through quantum corrections and becomes small around the electroweak energy scale. As a result, $`H_1^3`$ and $`H_2^3`$ can have non-vanishing v.e.v.s and assume the role of two Higgs doublets in the MSSM. For the first two generations, on the other hand, such couplings for $`H_1^i`$ or $`H_2^i`$ have small coefficients, so that the masses-squared of these scalar fields are kept around $`m_{3/2}^2`$. If the coefficient of $`S^3K^3K^{c3}`$ is large, the mass-squared of $`S^3`$ is also driven small. Although there is no phenomenological information about $`S^iK^iK^{ci}`$, a hierarchy of their coefficients could well exist. Among the three scalar fields $`S^i`$, one scalar field $`S^3`$ alone may have a non-vanishing v.e.v.. We thus assume that only $`H_1^3`$, $`H_2^3`$, and $`S^3`$ can have non-vanishing v.e.v.s. Quantum corrections to the masses-squared of other scalar fields are small, keeping them around $`m_{3/2}^2`$. The masses-squared of $`K^i`$, $`K^{ci}`$, and $`N^{ci}`$ receive non-negligible negative contributions from the D-term of U(1) when the gauge symmetry is broken spontaneously. However, the positive contributions from the supersymmetry-soft-breaking terms in Eq. (17) can dominate over and prevent these scalar fields from getting non-vanishing v.e.v.s. Assuming the above simplification, the scalar potential is given by $`V`$ $`=`$ $`{\displaystyle \frac{1}{8}}g_2^2\left(|H_1|^2+|H_2|^2\right)^2+{\displaystyle \frac{1}{8}}g_1^2\left(|H_1|^2|H_2|^2\right)^2`$ (21) $`+{\displaystyle \frac{1}{72}}g_{1}^{}{}_{}{}^{2}\left(4|H_1|^2+|H_2|^25|S|^2\right)^2`$ $`\left({\displaystyle \frac{1}{2}}g_2^2|\lambda _H|^2\right)|H_1H_2|^2+|\lambda _H|^2\left(|H_1|^2+|H_2|^2\right)|S|^2`$ $`+(B_H\lambda _Hm_{3/2}SH_1H_2+\mathrm{H}.\mathrm{c}.)+M_{H_1}^2|H_1|^2+M_{H_2}^2|H_2|^2+M_S^2|S|^2,`$ where the generation indices are left out. With group indices being expressed, $`H_1H_2`$ is written as $`ϵ_{ab}H_{1a}H_{2b}`$, so that holds an equation $`|H_1H_2|^2=|H_1|^2|H_2|^2|H_1^{}H_2|^2`$ . The gauge coupling constants for SU(2), U(1), and U(1) are denoted by $`g_2`$, $`g_1`$, and $`g_1^{}`$, respectively. We now discuss the v.e.v.s of the Higgs fields $`H_1`$, $`H_2`$, and $`S`$. For any values of $`H_1`$ and $`H_2`$, the complex phase of $`S`$ has a value which gives an equality $`B_H\lambda _Hm_{3/2}SH_1H_2=|B_H\lambda _Hm_{3/2}SH_1H_2|`$. For given values of $`|H_1|^2`$ and $`|H_2|^2`$, the v.e.v. $`|H_1H_2|^2`$ becomes maximum at $`H_1^{}H_2=0`$. Therefore, a condition $`g_2^2>2|\lambda _H|^2`$ guarantees electric charge conservation. Differently from the MSSM, there is no direction for the v.e.v.s where their quartic terms are absent in the scalar potential. The potential gives a stable vacuum irrespectively of the supersymmetry-soft-breaking terms. Redefining the global phases of the Higgs fields so as to give $`B_H\lambda _H=|B_H\lambda _H|`$, the v.e.v.s $`v_1`$, $`v_2`$, and $`v_s`$ of the neutral components of $`H_1`$, $`H_2`$, and $`S`$, respectively, may be taken real and non-negative. If these v.e.v.s are all non-vanishing, extremum conditions are given by $`{\displaystyle \frac{1}{8}}(g_2^2+g_1^2)(v_1^2v_2^2)v_1+{\displaystyle \frac{1}{18}}g_1^2(4v_1^2+v_2^25v_s^2)v_1`$ (22) $`+{\displaystyle \frac{1}{2}}|\lambda _H|^2(v_2^2+v_s^2)v_1{\displaystyle \frac{1}{\sqrt{2}}}|B_H\lambda _Hm_{3/2}|v_2v_s+M_{H_1}^2v_1=0,`$ (23) $`{\displaystyle \frac{1}{8}}(g_2^2+g_1^2)(v_1^2v_2^2)v_2+{\displaystyle \frac{1}{72}}g_1^2(4v_1^2+v_2^25v_s^2)v_2`$ (24) $`+{\displaystyle \frac{1}{2}}|\lambda _H|^2(v_1^2+v_s^2)v_2{\displaystyle \frac{1}{\sqrt{2}}}|B_H\lambda _Hm_{3/2}|v_1v_s+M_{H_2}^2v_2=0,`$ (25) $`{\displaystyle \frac{5}{72}}g_1^2(4v_1^2+v_2^25v_s^2)v_s+{\displaystyle \frac{1}{2}}|\lambda _H|^2(v_1^2+v_2^2)v_s`$ (26) $`{\displaystyle \frac{1}{\sqrt{2}}}|B_H\lambda _Hm_{3/2}|v_1v_2+M_S^2v_s=0.`$ (27) It turns out that the solution of these simultaneous equations is unique, if exists. The true vacuum is either at such a point or at a point where at least one v.e.v. vanishes, being determined by the potential energies of those points. The v.e.v.s of the Higgs bosons have to satisfy phenomenological constraints coming from experiments for the gauge bosons. The $`W`$-boson mass has been measured precisely. The $`Z`$ boson for SU(2)$`\times `$U(1) and the $`Z^{}`$ boson for U(1) are mixed and their mass-squared matrix is given by $`\left(\begin{array}{cc}M_Z^2& M_{ZZ^{}}^2\\ M_{ZZ^{}}^2& M_Z^{}^2\end{array}\right),`$ (28) $`M_Z^2`$ $`=`$ $`{\displaystyle \frac{1}{4}}(g_2^2+g_1^2)(v_1^2+v_2^2),`$ (29) $`M_Z^{}^2`$ $`=`$ $`{\displaystyle \frac{1}{36}}g_1^2(16v_1^2+v_2^2+25v_s^2),`$ (30) $`M_{ZZ^{}}^2`$ $`=`$ $`{\displaystyle \frac{1}{12}}g_1^{}\sqrt{g_2^2+g_1^2}(4v_1^2v_2^2).`$ (31) Two massive neutral gauge bosons, which are denoted by $`Z_1`$ and $`Z_2`$ ($`M_{Z_1}<M_{Z_2}`$), are predicted. The measured mass for the $`Z`$ boson of the SM should be taken as the mass of $`Z_1`$. The experimental lower bound on the mass of a new neutral gauge boson is about 600 GeV . According to detailed analyses of various experiments for an extra gauge boson , the mixing between $`Z`$ and $`Z^{}`$ is small. Defining a mixing parameter by $`R=(M_{ZZ^{}}^2)^2/M_Z^2M_Z^{}^2`$, a bound $`R\mathrm{\Gamma }<\mathrm{\hspace{0.17em}10}^3`$ is roughly obtained. The v.e.v.s can also be constrained by the lightest Higgs boson mass, whose experimental bound is given by $`M_{H^0}\mathrm{\Gamma }>\mathrm{\hspace{0.17em}80}`$ GeV . Since its predicted mass by the tree-level potential in Eq. (21) could be altered to become larger by several tens of GeV through one-loop quantum corrections, we conservatively put a constraint $`M_{H^0}>50`$ GeV to the tree-level mass. The scalar potential is analyzed numerically. For independent coefficients of the potential we choose $`|\lambda _H|`$, $`|B_H\lambda _Hm_{3/2}|`$, $`M_{H_1}^2`$, $`M_{H_2}^2`$, and $`M_S^2`$. In Fig. 1 we show the regions for $`M_{H_1}^2`$ and $`M_{H_2}^2`$ where the v.e.v.s are compatible with the above constraints. We have also imposed the constraints $`1v_2/v_135`$ and $`M_{Z_2}2000`$ GeV. With $`|B_H\lambda _Hm_{3/2}|`$ being 0.1 TeV, $`|\lambda _H|`$ is set for 0.1 and 0.3, which correspond to the upper and lower regions, respectively, For given values of $`M_{H_1}^2`$, $`M_{H_2}^2`$, $`|\lambda _H|`$, and $`|B_H\lambda _Hm_{3/2}|`$, the remaining parameter $`M_S^2`$ is so determined as to make the $`W`$-boson mass coincident with the measured value. The gauge coupling constant for U(1) is taken for $`g_1^{}=g_1`$. Owing to the constraints from $`M_{Z_2}`$ and $`R`$, in wide regions $`M_{H_1}^2`$ is larger than (1 TeV)<sup>2</sup>. The value of $`M_{H_2}^2`$ is generally smaller than $`M_{H_1}^2`$ in magnitude. The region for $`|\lambda _H|=0.3`$ with $`M_{H_1}^2\mathrm{\Gamma }<(500`$ GeV)<sup>2</sup> corresponds to $`v_2/v_1\mathrm{\Gamma }<\mathrm{\hspace{0.17em}2}`$. A rough estimate of Eq. (25) shows that the sign of $`M_{H_2}^2`$ is positive for $`|\lambda _H|^2<(5/36)g_1^2`$ while $`M_{H_2}^2`$ has either sign for larger values of $`|\lambda _H|`$. The value of $`M_S^2`$ is smaller than $`M_{H_2}^2`$ and always negative. As $`|B_H\lambda _Hm_{3/2}|`$ increases, the allowed values for $`M_{H_1}^2`$ become larger, which is seen from Eq. (23). If the upper limit for $`M_{Z_2}`$ is lifted, wider parameter regions become allowed. However, as the scale of the mass-squared parameters increases, more fine-tuning of the parameters becomes inevitable for electroweak symmetry breaking. For having the correct vacuum, large differences among $`M_{H_1}^2`$, $`M_{H_2}^2`$, and $`M_S^2`$ are necessary, which could well occur under our assumption for supersymmetry breaking. In Table III we present four examples for the values of $`M_{H_1}^2`$, $`M_{H_2}^2`$, and $`M_S^2`$ in the allowed regions of Fig. 1. Also shown are the resultant values for $`v_2/v_1`$, $`v_s`$, $`M_{Z_2}`$, $`R`$, and the masses of the physical Higgs bosons. These Higgs-boson masses have been calculated, assuming for definiteness that the mass eigenstates are formed by the Higgs fields $`H_1`$, $`H_2`$, and $`S`$ without mixing with the other fields of $`H_1^i`$, $`H_2^i`$, and $`S^i`$. Therefore, there are three mass eigenstates for the neutral scalar bosons $`H^0`$, one for the neutral pseudoscalar boson $`A^0`$, and one for the charged scalar boson $`H^\pm `$. One neutral scalar boson is light, whereas the others have large masses. The mixing parameter $`R`$ vanishes for $`v_2/v_1=2`$, as seen from Eq. (31). The large mass difference between $`Z_1`$ and $`Z_2`$ requires in some degree fine-tuning for the parameters in the potential. Since these two masses are different from each other by one order of magnitude, it is generally necessary to adjust the values of the mass-squared parameters $`M_{H_1}^2`$, $`M_{H_2}^2`$, $`M_S^2`$ and the coupling constants $`\lambda _H`$, $`B_H`$ within the accuracy of order $`10^2`$. In Fig. 2, for the four examples in Table III, the ratio of a predicted $`W`$-boson mass to the experimental value is depicted as a function of $`M_S^2`$ normalized to its proper value which yields the correct $`W`$-boson mass. If the value of one parameter alone in the potential is deviated by order of $`10^1`$, the resultant v.e.v.s lead to a $`W`$-boson mass different from its experimental value by a factor or more. The neutrinos have both Dirac and Majorana masses. Neglecting the generation mixing, the mass matrix for the left-handed and right-handed neutrinos becomes $$\left(\begin{array}{cc}0& \eta _\nu v_2/\sqrt{2}\\ \eta _\nu v_2/\sqrt{2}& \sqrt{2}\lambda _Nv_s\end{array}\right),$$ (32) whose lighter mass eigenvalue is approximately given by $`m_{\nu 1}=|\eta _\nu |^2v_2^2/2\sqrt{2}|\lambda _N|v_s`$. With $`|\lambda _N|=0.2`$ and $`v_s=3`$ TeV, $`m_{\nu 1}`$ becomes about 59 eV for $`|\eta _\nu |v_2=10`$ MeV and 0.59 eV for $`|\eta _\nu |v_2=1`$ MeV. Even if the Yukawa coupling constants for the neutrino Dirac masses are of the same order as that for the electron, the observed ordinary neutrinos have tiny masses which could be a reason for the recent experimental results suggesting neutrino oscillations. The coupling $`SH_1H_2`$ serves as the $`\mu `$ term and an effective $`\mu `$ parameter is given by $`\mu =\lambda _Hv_s/\sqrt{2}`$. For $`|\lambda _H|=0.2`$ and $`v_s=3`$ TeV, $`|\mu |`$ is approximately 420 GeV. The parameter $`\mu `$ can have an appropriate magnitude for the electroweak symmetry breaking. This parameter also affects the masses of the charginos and the neutralinos. Assuming that these particles are formed by the fermion components of $`H_1`$, $`H_2`$, and $`S`$, as well as the SU(2), U(1), and U(1) gauge fermions, there exist two mass eigenstates for the charginos and six mass eigenstates for the neutralinos. Provided that the gauge fermions for SU(2), U(1), and U(1) receive masses of order 100 GeV from the supersymmetry-soft-breaking terms in Eq. (17), the masses of the lighter chargino and the lightest neutralino become of order 100 GeV. This model contains new particles which are not predicted by the MSSM. As already noted, the gauge-Higgs sector involves extra a neutral gauge boson, a neutral Higgs boson, and two neutralinos. For the lepton sector, there appear a heavy neutrino in each generation. Correspondingly the scalar neutrinos are duplicated. The interactions arising from the coupling $`SN^cN^c`$ do not conserve lepton number. In addition, the superfields $`H_1^i`$, $`H_2^i`$, $`S^i`$ with $`i=1,2`$ and $`K^j`$, $`K^{cj}`$ with $`j=1,2,3`$ are newly introduced. The masses of their fermion components are generated by the couplings to $`H_1^3`$, $`H_2^3`$, and $`S^3`$, and become of order $`0.11`$ TeV. The lightest fermion among $`K^j`$ and $`K^{cj}`$ is stable. As well as by collider experiments, such a stable particle may be explored by other methods to search for its relics in the universe, e.g. anomalous nuclei in sea water. However, these methods depend on the relic density, whose theoretical prediction is plagued by various uncertainties for non-perturbative effects, cosmology, and so on. Since the scalar components of $`H_1^i`$ and $`H_2^i`$ couple to quarks and leptons, non-trivial constraints are imposed on their coupling constants from the viewpoint of flavor-changing neutral current. However, these scalar particles are rather heavy, so that the constraints are not so stringent as usually thought. If the couplings $`S^iH_1^3H_2^3`$, $`S^3H_1^iH_2^3`$, or $`S^3H_1^3H_2^i`$ are not neglected, some or all of the scalar or fermion components for $`H_1^i`$, $`H_2^i`$, and $`S^i`$ are mixed with the Higgs bosons, charginos, or neutralinos, leading to an enlargement of the particles belonging to the gauge-Higgs sector. The MSSM has another problem on the EDMs of the neutron and the electron, which can be explained in this model. If the squark and slepton masses are of order 100 GeV and the $`CP`$-violating phases intrinsic in the model are not suppressed, these EDMs are predicted to be much larger than their experimental upper bounds. However, a typical scale of the squark and slepton masses in this model, which are considered not much different from the mass of $`H_1`$, is larger than 1 TeV. The EDMs then lie within the experimental bounds without fine-tuning the $`CP`$-violating phases to be very small . If these phases are not suppressed, the interactions of the charginos or the neutralinos generally induce sizable $`CP`$ violation in their production or decay processes. ## IV Energy Dependence The parameter values of the model change according to the relevant energy scale. Analyzing their energy dependencies, we discuss whether gauge symmetry breaking is induced by radiative corrections. We also examine the scenario of universal values for the masses-squared and the trilinear coupling constants of scalar fields at a very high energy scale. For simplicity, the generation mixing of the particle fields are neglected. The evolution of the parameters concerning the energy-scale change are described by RGEs, which are given in Appendix A. It is seen from those equations that $`M_{H_2}^2`$ increases as the energy scale becomes high, owing to a large Yukawa coupling constant $`\eta _u`$ for the top quark. If the coupling constant $`\lambda _K`$ is around unity, $`M_S^2`$ also increases. Consequently, the mass-squared parameters can all have large positive values at high energy scales, even if they are small at a low energy scale as discussed in the previous section. The SU(2)$`\times `$U(1)$`\times `$U(1) symmetry is spontaneously broken through radiative corrections. The experimental values of the gauge coupling constants suggest that these constants are not unified at the energy scale for possible grand unification. This gauge unification could be achieved by incorporating one additional pair of SU(2)-doublet chiral superfields. However, such a pair form a gauge-singlet linear coupling and thus ruin the model by necessitating a mass parameter of unknown origin. Although the particle contents are not embedded in the fundamental representation of the E<sub>6</sub> group, the masses and the coupling constants evolve similarly to those in the E<sub>6</sub> models. Some features of these models apply to the present model, and vice versa. We now numerically examine the evolution of the parameters. Taking the masses-squared and the trilinear coupling constants of the scalar fields for common values $`m_{3/2}^2`$ and $`A`$ at a high energy scale $`M_X`$, we evaluate the parameters at a low energy scale $`M`$. For definiteness, we set $`M_X`$ for $`10^{17}`$ GeV and $`M`$ for $`5\times 10^2`$ GeV. Assuming an equality $`g_1=g_1^{}`$, all the gauge coupling constants are determined independently of the parameters. The masses of the gauge fermions are also determined, if their values are given at some energy scale. Since the gauge groups are not unified in our model, these masses at $`M_X`$ are generally different from each other. However, they are considered nevertheless to be of the same order of magnitude. We therefore put their values equal at $`M_X`$, $`\stackrel{~}{m}_3=\stackrel{~}{m}_2=\stackrel{~}{m}_1=\stackrel{~}{m}_1^{}\stackrel{~}{m}`$, for simplicity. The Yukawa coupling constants $`\eta _u`$, $`\eta _d`$, $`\eta _\nu `$, $`\eta _e`$, $`\lambda _N`$, $`\lambda _H`$, and $`\lambda _K`$ at $`M_X`$, which are specified by attaching an index ’$`X`$’, are independent of each other. In Fig. 3 the values of $`\eta _u`$, $`\lambda _N`$, $`\lambda _H`$, and $`\lambda _K`$ at $`M`$ are depicted as functions of $`\eta _u^X`$. We have taken $`\lambda _N^X=\lambda _H^X=\lambda _K^X=0.2`$ and $`\eta _d^X=\eta _\nu ^X=\eta _e^X=0`$. The magnitude of $`\eta _u`$ and $`\lambda _K`$ become large at the low energy scale, while the energy dependencies of $`\lambda _N`$ and $`\lambda _H`$ are not significant. The evolution of each Yukawa coupling constant is not affected much by the other parameters. For generating an appropriate mass for the top quark, $`|\eta _u^X|`$ should be larger than 0.1. The condition $`g_2^2>2|\lambda _H|^2`$ for electric charge conservation at the low energy scale is satisfied for $`|\lambda _H^X|\mathrm{\Gamma }<\mathrm{\hspace{0.17em}0.5}`$. In Fig. 4 we show the trilinear coupling constant $`B_H`$ as a function of $`A`$ for three sets of $`\eta _u^X`$, $`\lambda _K^X`$, and $`\stackrel{~}{m}/m_{3/2}`$ given in Table IV. For the other non-vanishing input parameters, we take $`\lambda _N^X=\lambda _H^X=0.2`$. The magnitude of $`A`$ is constrained as $`|A|<3`$ in order not to induce incorrect breaking of gauge symmetry. The value of $`B_H`$ is of order unity in wide ranges of $`A`$ except narrow ranges where $`|B_H|`$ is much smaller than unity. The $`\stackrel{~}{m}`$ dependence of $`B_H`$ is negligible. The mass-squared parameters $`M_{H_1}^2`$, $`M_{H_2}^2`$, and $`M_S^2`$ are shown, as functions of $`A`$, for the three parameter sets (a), (b), and (c) of Table IV in Figs. 5(a), 5(b), and 5(c), respectively, with $`\lambda _N^X=\lambda _H^X=0.2`$. The gravitino mass is fixed as $`m_{3/2}=2`$ TeV. At the low energy scale, receiving large quantum corrections, $`M_{H_2}^2`$ and $`M_S^2`$ could become much smaller than the universal value $`m_{3/2}^2`$. These corrections strongly depend on $`\eta _u^X`$, $`\lambda _K^X`$, $`\stackrel{~}{m}`$, and $`A`$. In Fig. 5(a), $`M_{H_2}^2`$ and $`M_S^2`$ become negative only for $`|A|\mathrm{\Gamma }>\mathrm{\hspace{0.17em}2}`$. However, as $`\eta _u^X`$ and $`\lambda _K^X`$ increase, $`M_{H_2}^2`$ and $`M_S^2`$ decrease, respectively. These masses-squared also become small for larger values of $`\stackrel{~}{m}`$. In Figs. 5(b) and 5(c), $`M_{H_2}^2`$ and $`M_S^2`$ are negative for any value of $`A`$. If $`\lambda _K^X`$ is larger than $`\eta _u^X`$, an inequality $`M_S^2<M_{H_2}^2`$ holds at the low energy scale. On the other hand, $`M_{H_1}^2`$ is not much different from $`m_{3/2}^2`$. We see from Figs. 3, 4, and 5 that the gauge symmetry of the vacuum at high energy scales can be spontaneously broken at low energy scales through radiative corrections. Furthermore, certain parameter values at a high energy scale, with the masses-squared and the trilinear coupling constants of scalar fields being universal, lead to the values of $`M_{H_1}^2`$, $`M_{H_2}^2`$, $`M_S^2`$, $`\lambda _H`$, and $`B_H`$ which give a plausible vacuum around the electroweak energy scale. As explicit examples, we show in Table V the parameter values at $`M_X`$ which give low-energy vacua consistent with experimental results. Since there are many input parameters, for simplicity, we have fixed the trilinear coupling constant and the Yukawa coupling constants as $`A=1`$ and $`\eta _u^X=\lambda _N^X=\lambda _K^X=0.2`$. The values of $`m_{3/2}`$ and $`\stackrel{~}{m}`$ are set for $`m_{3/2}=1`$, 2 TeV and $`\stackrel{~}{m}=0.3`$, 0.5 TeV. The resultant values of $`v_2/v_1`$, $`v_s`$, $`M_{Z_2}`$, $`R`$, and the Higgs boson masses at $`M`$ are also given together with the masses of the $`W`$ boson $`M_W`$, the top quark $`m_t`$, and the bottom quark $`m_b`$. This model is compatible with the supersymmetry breaking mechanism based on $`N=1`$ supergravity. ## V Conclusions The MSSM involves the problems on proton lifetime, neutrino masses, and the $`\mu `$ term. These problems may be solved by theories at very high energy scales, which however are mere conjectures and mostly untestable at present. On the contrary, the solutions of the problems may reside in theories around an energy scale of the MSSM or the SM. Then, a new model beyond the MSSM should exists. Such a model is rigidly constructed on solid ground of various experimental results. Its predictions can be examined at experiments in the near future. To solve the problems of the MSSM without invoking uncertain theories, we have constructed a supersymmetric standard model based on SU(3)$`\times `$SU(2)$`\times `$U(1)$`\times `$U(1) gauge symmetry and $`N=1`$ supergravity. In this model, the solutions are given consistently within the framework of the model. The interactions of dimension four or five which violate baryon number conservation are not allowed by the gauge symmetry, leading to an adequately long lifetime of the proton. The gauge symmetry also prescribes the existence of right-handed neutrinos and an SU(2)-singlet Higgs boson whose v.e.v. $`v_s`$ induces breaking of the U(1) symmetry. The neutrino masses receive contributions of Dirac type and those of Majorana type generated by the large v.e.v. $`v_s`$. The ordinary neutrinos then have tiny masses. The $`\mu `$ term of the MSSM is replaced by a trilinear coupling of the Higgs superfields, and the effective $`\mu `$ parameter is given by $`v_s`$. A typical mass scale of this model is of order 1 TeV. Some fine-tuning of the parameters is necessary for correct breaking of the electroweak gauge symmetry. On the other hand, the EDMs of the neutron and the electron are predicted within their experimental bounds without fine-tuning much $`CP`$-violating phases. The constructed model gives predictions different from the MSSM in various phenomenological aspects. An extra neutral gauge boson couples to all the quarks and leptons. There exists a stable fermion which is nontrivially transformed under SU(3) and U<sub>EM</sub>(1). Lepton number is not conserved in the interactions of the neutrinos or the scalar neutrinos. Some scalar particles mediate flavor-changing neutral current at the tree level. The experimental examination of these predictions could be performed in the near future. Implications of this model for theories at higher energy scales have also been studied. The gauge symmetry breaking is induced by radiative corrections. The masses-squared and the trilinear coupling constants of scalar fields could be universal at an energy scale not much smaller than the Planck mass, which is consistent with the mechanism of supersymmetry breaking based on $`N=1`$ supergravity. The gauge coupling constants are not unified below the Planck mass scale, unless the particle contents are modified. ###### Acknowledgements. We thank T. Watanabe for his advice on computer calculations. One of us (M.A.) acknowledges the Japan Society for the Promotion of Science for financial support. The work of M.A. is supported in part by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture, Japan. ## A The RGEs are listed below. The gauge and Yukawa coupling constants are expressed by $`\alpha g^2/4\pi `$ and $`E\eta ^2/4\pi `$, $`L\lambda ^2/4\pi `$. The gauge coupling constants and the gauge fermion masses: $`\mu {\displaystyle \frac{d}{d\mu }}\alpha _a`$ $`=`$ $`{\displaystyle \frac{b_a}{2\pi }}\alpha _a^2,`$ (33) $`\mu {\displaystyle \frac{d}{d\mu }}\stackrel{~}{m}_a`$ $`=`$ $`{\displaystyle \frac{b_a}{2\pi }}\alpha _a\stackrel{~}{m}_a,`$ (34) where $`b_3=0`$, $`b_2=3`$, $`b_1=15`$, and $`b_1^{}=15`$ for SU(3), SU(2), U(1), and U(1), respectively. The masses-squared of the scalar fields: $`\mu {\displaystyle \frac{d}{d\mu }}M_Q^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3^2+{\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2^2+{\displaystyle \frac{1}{36}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{1}{144}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{6}}\alpha _1\xi +{\displaystyle \frac{1}{12}}\alpha _1^{}\xi ^{}\right)`$ (37) $`+{\displaystyle \frac{1}{2\pi }}E_u\left(|A_u|^2m_{3/2}^2+M_{H_2}^2+M_Q^2+M_{U^c}^2\right)`$ $`+{\displaystyle \frac{1}{2\pi }}E_d\left(|A_d|^2m_{3/2}^2+M_{H_1}^2+M_Q^2+M_{D^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{U^c}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3^2+{\displaystyle \frac{4}{9}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{1}{144}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{2}{3}}\alpha _1\xi +{\displaystyle \frac{1}{12}}\alpha _1^{}\xi ^{}\right)`$ (39) $`+{\displaystyle \frac{1}{\pi }}E_u\left(|A_u|^2m_{3/2}^2+M_{H_2}^2+M_Q^2+M_{U^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{D^c}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3^2+{\displaystyle \frac{1}{9}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{49}{144}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{3}}\alpha _1\xi +{\displaystyle \frac{7}{12}}\alpha _1^{}\xi ^{}\right)`$ (41) $`+{\displaystyle \frac{1}{\pi }}E_d\left(|A_d|^2m_{3/2}^2+M_{H_1}^2+M_Q^2+M_{D^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_L^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2^2+{\displaystyle \frac{1}{4}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{49}{144}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{2}}\alpha _1\xi +{\displaystyle \frac{7}{12}}\alpha _1^{}\xi ^{}\right)`$ (44) $`+{\displaystyle \frac{1}{2\pi }}E_\nu \left(|A_\nu |^2m_{3/2}^2+M_{H_2}^2+M_L^2+M_{N^c}^2\right)`$ $`+{\displaystyle \frac{1}{2\pi }}E_e\left(|A_e|^2m_{3/2}^2+M_{H_1}^2+M_L^2+M_{E^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{N^c}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{25}{144}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{5}{12}}\alpha _1^{}\xi ^{}\right)`$ (47) $`+{\displaystyle \frac{1}{\pi }}E_\nu \left(|A_\nu |^2m_{3/2}^2+M_{H_2}^2+M_L^2+M_{N^c}^2\right)`$ $`+{\displaystyle \frac{2}{\pi }}L_N\left(|B_N|^2m_{3/2}^2+M_S^2+2M_{N^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{E^c}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left(\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{1}{144}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left(\alpha _1\xi +{\displaystyle \frac{1}{12}}\alpha _1^{}\xi ^{}\right)`$ (49) $`+{\displaystyle \frac{1}{\pi }}E_e\left(|A_e|^2m_{3/2}^2+M_{H_1}^2+M_L^2+M_{E^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{H_1}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2^2+{\displaystyle \frac{1}{4}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{4}{9}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{2}}\alpha _1\xi {\displaystyle \frac{2}{3}}\alpha _1^{}\xi ^{}\right)`$ (53) $`+{\displaystyle \frac{3}{2\pi }}E_d\left(|A_d|^2m_{3/2}^2+M_{H_1}^2+M_Q^2+M_{D^c}^2\right)`$ $`+{\displaystyle \frac{1}{2\pi }}E_e\left(|A_e|^2m_{3/2}^2+M_{H_1}^2+M_L^2+M_{E^c}^2\right)`$ $`+{\displaystyle \frac{1}{2\pi }}L_H\left(|B_H|^2m_{3/2}^2+M_S^2+M_{H_1}^2+M_{H_2}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{H_2}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2^2+{\displaystyle \frac{1}{4}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{1}{36}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{2}}\alpha _1\xi {\displaystyle \frac{1}{6}}\alpha _1^{}\xi ^{}\right)`$ (57) $`+{\displaystyle \frac{3}{2\pi }}E_u\left(|A_u|^2m_{3/2}^2+M_{H_2}^2+M_Q^2+M_{U^c}^2\right)`$ $`+{\displaystyle \frac{1}{2\pi }}E_\nu \left(|A_\nu |^2m_{3/2}^2+M_{H_2}^2+M_L^2+M_{N^c}^2\right)`$ $`+{\displaystyle \frac{1}{2\pi }}L_H\left(|B_H|^2m_{3/2}^2+M_S^2+M_{H_1}^2+M_{H_2}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_S^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{25}{36}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{5}{6}}\alpha _1^{}\xi ^{}\right)`$ (61) $`+{\displaystyle \frac{1}{\pi }}L_N\left(|B_N|^2m_{3/2}^2+M_S^2+2M_{N^c}^2\right)`$ $`+{\displaystyle \frac{1}{\pi }}L_H\left(|B_H|^2m_{3/2}^2+M_S^2+M_{H_1}^2+M_{H_2}^2\right)`$ $`+{\displaystyle \frac{3}{2\pi }}L_K\left(|B_K|^2m_{3/2}^2+M_S^2+M_K^2+M_{K^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_K^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3^2+{\displaystyle \frac{1}{9}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{4}{9}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{3}}\alpha _1\xi {\displaystyle \frac{2}{3}}\alpha _1^{}\xi ^{}\right)`$ (63) $`+{\displaystyle \frac{1}{2\pi }}L_K\left(|B_K|^2m_{3/2}^2+M_S^2+M_K^2+M_{K^c}^2\right),`$ $`\mu {\displaystyle \frac{d}{d\mu }}M_{K^c}^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3^2+{\displaystyle \frac{1}{9}}\alpha _1\stackrel{~}{m}_1^2+{\displaystyle \frac{1}{36}}\alpha _1^{}\stackrel{~}{m}_1^2\right)+{\displaystyle \frac{1}{2\pi }}\left({\displaystyle \frac{1}{3}}\alpha _1\xi {\displaystyle \frac{1}{6}}\alpha _1^{}\xi ^{}\right)`$ (65) $`+{\displaystyle \frac{1}{2\pi }}L_K\left(|B_K|^2m_{3/2}^2+M_S^2+M_K^2+M_{K^c}^2\right),`$ where $`\xi =Y_\varphi M_\varphi ^2`$ and $`\xi ^{}=Q_\varphi M_\varphi ^2`$. The Yukawa coupling constants: $`\mu {\displaystyle \frac{d}{d\mu }}E_u`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}E_u\left\{{\displaystyle \frac{4}{3}}\alpha _3+{\displaystyle \frac{3}{4}}\alpha _2+{\displaystyle \frac{13}{36}}\alpha _1+{\displaystyle \frac{1}{48}}\alpha _1^{}{\displaystyle \frac{1}{4}}(6E_u+E_d+E_\nu +L_H)\right\},`$ (66) $`\mu {\displaystyle \frac{d}{d\mu }}E_d`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}E_d\left\{{\displaystyle \frac{4}{3}}\alpha _3+{\displaystyle \frac{3}{4}}\alpha _2+{\displaystyle \frac{7}{36}}\alpha _1+{\displaystyle \frac{19}{48}}\alpha _1^{}{\displaystyle \frac{1}{4}}(E_u+6E_d+E_e+L_H)\right\},`$ (67) $`\mu {\displaystyle \frac{d}{d\mu }}E_\nu `$ $`=`$ $`{\displaystyle \frac{2}{\pi }}E_\nu \left\{{\displaystyle \frac{3}{4}}\alpha _2+{\displaystyle \frac{1}{4}}\alpha _1+{\displaystyle \frac{13}{48}}\alpha _1^{}{\displaystyle \frac{1}{4}}(3E_u+4E_\nu +E_e+4L_N+L_H)\right\},`$ (68) $`\mu {\displaystyle \frac{d}{d\mu }}E_e`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}E_e\left\{{\displaystyle \frac{3}{4}}\alpha _2+{\displaystyle \frac{3}{4}}\alpha _1+{\displaystyle \frac{19}{48}}\alpha _1^{}{\displaystyle \frac{1}{4}}(3E_d+E_\nu +4E_e+L_H)\right\},`$ (69) $`\mu {\displaystyle \frac{d}{d\mu }}L_N`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}L_N\left\{{\displaystyle \frac{25}{48}}\alpha _1^{}{\displaystyle \frac{1}{4}}(4E_\nu +10L_N+2L_H+3L_K)\right\},`$ (70) $`\mu {\displaystyle \frac{d}{d\mu }}L_H`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}L_H\{{\displaystyle \frac{3}{4}}\alpha _2+{\displaystyle \frac{1}{4}}\alpha _1+{\displaystyle \frac{7}{12}}\alpha _1^{}`$ (72) $`{\displaystyle \frac{1}{4}}(3E_u+3E_d+E_\nu +E_e+2L_N+4L_H+3L_K)\},`$ $`\mu {\displaystyle \frac{d}{d\mu }}L_K`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}L_K\left\{{\displaystyle \frac{4}{3}}\alpha _3+{\displaystyle \frac{1}{9}}\alpha _1+{\displaystyle \frac{7}{12}}\alpha _1^{}{\displaystyle \frac{1}{4}}(2L_N+2L_H+5L_K)\right\}.`$ (73) The trilinear coupling constants: $`\mu {\displaystyle \frac{d}{d\mu }}A_u`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3+{\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2+{\displaystyle \frac{13}{36}}\alpha _1\stackrel{~}{m}_1+{\displaystyle \frac{1}{48}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (75) $`+{\displaystyle \frac{1}{2\pi }}(6A_uE_u+A_dE_d+A_\nu E_\nu +B_HL_H),`$ $`\mu {\displaystyle \frac{d}{d\mu }}A_d`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3+{\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2+{\displaystyle \frac{7}{36}}\alpha _1\stackrel{~}{m}_1+{\displaystyle \frac{19}{48}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (77) $`+{\displaystyle \frac{1}{2\pi }}(A_uE_u+6A_dE_d+A_eE_e+B_HL_H),`$ $`\mu {\displaystyle \frac{d}{d\mu }}A_\nu `$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2+{\displaystyle \frac{1}{4}}\alpha _1\stackrel{~}{m}_1+{\displaystyle \frac{13}{48}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (79) $`+{\displaystyle \frac{1}{2\pi }}(3A_uE_u+4A_\nu E_\nu +A_eE_e+4B_NL_N+B_HL_H),`$ $`\mu {\displaystyle \frac{d}{d\mu }}A_e`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2+{\displaystyle \frac{3}{4}}\alpha _1\stackrel{~}{m}_1+{\displaystyle \frac{19}{48}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (81) $`+{\displaystyle \frac{1}{2\pi }}(3A_dE_d+A_\nu E_\nu +4A_eE_e+B_HL_H),`$ $`\mu {\displaystyle \frac{d}{d\mu }}B_N`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{25}{48}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (83) $`+{\displaystyle \frac{1}{2\pi }}(4A_\nu E_\nu +10B_NL_N+2B_HL_H+3B_KL_K),`$ $`\mu {\displaystyle \frac{d}{d\mu }}B_H`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{3}{4}}\alpha _2\stackrel{~}{m}_2+{\displaystyle \frac{1}{4}}\alpha _1\stackrel{~}{m}_1+{\displaystyle \frac{7}{12}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (85) $`+{\displaystyle \frac{1}{2\pi }}(3A_uE_u+3A_dE_d+A_\nu E_\nu +A_eE_e+2B_NL_N+4B_HL_H+3B_KL_K),`$ $`\mu {\displaystyle \frac{d}{d\mu }}B_K`$ $`=`$ $`{\displaystyle \frac{2}{\pi }}\left({\displaystyle \frac{4}{3}}\alpha _3\stackrel{~}{m}_3+{\displaystyle \frac{1}{9}}\alpha _1\stackrel{~}{m}_1+{\displaystyle \frac{7}{12}}\alpha _1^{}\stackrel{~}{m}_1^{}\right){\displaystyle \frac{1}{m_{3/2}}}`$ (87) $`+{\displaystyle \frac{1}{2\pi }}(2B_NL_N+2B_HL_H+5B_KL_K).`$
warning/0003/cond-mat0003344.html
ar5iv
text
# Impurity-induced dephasing of Andreev states ## I Introduction How much can flicker noise in the junction transparency affect coherent transport in Andreev states present in a superconducting quantum point contact (SQPC)? The assumption of coherent transport in Andreev states is widely used in theoretical work, see e. g. Refs. and . However, in realistic systems interaction with a dynamical environment will always introduce some amount of dephasing, see Refs. and for a review. In the so-called microwave-activated quantum interferometer full coherency of Andreev states is assumed and a method for Andreev-level spectroscopy is presented. The Andreev levels are probed with a microwave field, resulting in an interference pattern in the current. If dephasing is present this interference pattern will begin to deteriorate. This connection between dephasing and current makes the microwave-activated quantum interferometer a suitable system to study the effect of flicker noise in the junction transparency on transport through Andreev states. It also provides an excellent opportunity to probe the coupling between a SQPC and its electromagnetic environment. In the following we study the role of dephasing induced by flicker noise in the normal state junction transparency, $`D`$, of a SQPC in the transport through Andreev states. The concrete system which will be considered is the above presented interferometer. Flicker noise can be caused by the presence of an impurity atom close to the junction which has two states of almost equal energy to choose from. When the atom tunnels between its two states the junction transparency will fluctuate. Another source of flicker noise is the tunneling of an electron between impurity atoms in a doped region. If there are two such neighboring defects with available states, a hybrid two-level state is formed and the electron can hop between the two. This hopping will then add a fluctuation to the junction transparency. The amplitude of these fluctuations depend on the distance between the defects and the junction. From now on we will refer to these dynamic defects as two-level elementary fluctuators (EFs). The microwave-activated quantum interferometer (MAQI) is based on a short, weakly biased SQPC which is subject to microwave irradiation. Confined to the contact area there are current carrying Andreev states. The corresponding energy levels – Andreev levels – are found in pairs within the superconductor energy gap $`\mathrm{\Delta }`$, one below and one above the Fermi level. When a SQPC is short ($`L\xi _0`$ where $`L`$ is the length of the junction while $`\xi _0`$ is the superconductor coherence length), there is only one pair of Andreev levels and their positions depend on the order parameter phase difference, $`\varphi `$, across the contact as $$E_\pm =\pm E(\varphi )=\pm \mathrm{\Delta }\sqrt{1D\mathrm{sin}^2(\varphi /2)}.$$ (1) Within this pair, the two states carry current in opposite directions and in equilibrium only the lower state is populated. The applied bias, $`V`$, through the Josephson relation $`\dot{\varphi }=2eV/\mathrm{}`$, forces the Andreev levels to move adiabatically within the energy gap with a period of $`T_p=\mathrm{}\pi /eV`$, see Fig. 1. The microwave field induces Landau-Zener (LZ) transitions between the Andreev levels (symbolized by wavy lines in Fig. 1). If the upper level is populated after the second transition a delocalized quasiparticle excitation will be created when this Andreev level merges with the continuum. The result will be a dc contribution to the current. Further, this current exhibits an interference pattern since there are two paths available to the upper level. It is this “interference effect” which is utilized in the MAQI for Andreev-level spectroscopy. In order to estimate the influence of flicker noise on the interference pattern we continue with a quantitative presentation of the interferometer and the model used for the fluctuations. ## II Model and method Consider a short single-mode SQPC which is subject to a high frequency microwave field ($`\mathrm{}\omega =2E(\varphi )2\mathrm{\Delta }`$). Let the contact, placed at $`x=0`$, be characterized by an energy-independent transparency $`D`$. A weak bias, $`eV\mathrm{\Delta }`$, is applied across the junction. We choose to describe the quasiparticles in the contact region with the following wave function, $$\mathrm{\Psi }(x,t)=u_+(x,t)e^{ik_Fx}+u_{}(x,t)e^{ik_Fx},$$ (2) where the envelope functions $`u_\pm (x,t)`$, left and right movers, are two-component vectors in electron-hole space. To simplify notation we introduce the four-component vector $`𝐮=[u_+,u_{}]`$. This vector satisfies the time-dependent Bogoliubov-de Gennes equation $`i\mathrm{}𝐮/t`$ $`=`$ $`[_0+\sigma _zV_g(t)]𝐮,`$ (3) $`_0`$ $`=`$ $`i\mathrm{}v_F\sigma _z\tau _z/x+\mathrm{\Delta }[\sigma _x\mathrm{cos}(\varphi (t)/2)`$ (5) $`+\text{sgn}(x)\sigma _y\mathrm{sin}(\varphi (t)/2)],`$ where $`\sigma _i`$ and $`\tau _i`$ denote Pauli matrices in electron-hole space and in $`\pm `$ space, respectively, while $`V_g(t)=V_\omega \mathrm{cos}(\omega t)`$ is the time-dependent gate potential. We assume $`eV_\omega \mathrm{\Delta }`$. The boundary condition at $`x=0`$ is $$𝐮(+0)=D^{1/2}\left[1\tau _y(1D)^{1/2}\right]𝐮(0).$$ As a result of the applied high frequency field there will be resonant transitions between the Andreev levels. These transitions introduce a mixed state which can be described within the resonance approximation as $$𝐮(x,t)=\underset{\pm }{}b^\pm (t)𝐮^\pm (x)e^{\omega t/2},$$ where $`𝐮^+(x)`$ and $`𝐮^{}(x)`$ are the envelope functions of the upper and lower Andreev states, while $`b^+`$ and $`b^{}`$ are the corresponding probability amplitudes. The final result is a dc current through the SQPC, $`I_{\mathrm{dc}}`$ $`=`$ $`{\displaystyle \frac{e(2\mathrm{\Delta }\mathrm{}\omega )}{\mathrm{}\pi }}\left|b^+\left({\displaystyle \frac{\mathrm{}\pi }{eV}}\right)\right|^2=2I_0\mathrm{sin}^2(\mathrm{\Theta }+\mathrm{\Phi }),`$ (6) $`I_0`$ $``$ $`(2e/\mathrm{}\pi )r^2(1r^2)\left(2\mathrm{\Delta }\mathrm{}\omega \right),`$ (7) where $`r^2`$ is the LZ transition amplitude, which depends on the bias voltage and the amplitude of the perturbation ($`\mathrm{\Theta }`$ is the phase of the LZ transition, which can be considered constant). The phase $`\mathrm{\Phi }`$ which inhibits the interference is calculated through $$\mathrm{\Phi }=\frac{1}{2eV}_{\varphi _A}^{\varphi _B}\left(E(\varphi )\frac{\mathrm{}\omega }{2}\right)d\varphi .$$ (8) Flicker noise of the junction transparency, $`D`$, which is the subject to be discussed now, enters through $`E(\varphi )`$, Eq. (1), in this expression. ### A Flicker noise in the junction transparency The main topic of this work is to study the effect of fluctuations in the junction transparency on the MAQI. For simplicity we choose to model the sources of these fluctuations, the EFs, with the so-called *random telegraph process*. This process is characterized by a random quantity $`\xi (t)`$ which has the value $`+1`$ or $`1`$ depending on whether the upper or lower EF state is occupied. We assume that the probability of each state is the same, namely $`1/2`$. This is acceptable since EFs with inter-level distances, $`E_i`$, smaller than $`k_\text{B}T`$, will be “frozen” — they behave as static impurities which do not affect the dynamic fluctuations of $`D`$. In this model the EF switches between its two states randomly in time. Physically, switching is a result of interactions between the EF and phonons or electrons in the contact area. In the presence of EFs the junction transparency will be modulated. In other words, $`DD+D_f(t)`$, where $`D_f(t)`$ is assumed to be small. Generally $$D_f(t)=\underset{i}{}A_i\xi _i(t),$$ with $`A_i`$ being the coupling strength of the $`i`$-th EF. We assume that the random processes in different EFs are not correlated. Consequently, after a change in variables from $`t`$ to $`\varphi `$ we can specify the random telegraph processes $`\xi _i(t)`$ through the correlation function, $$\xi _i(\varphi _1)\xi _j(\varphi _2)=\delta _{ij}e^{2\gamma _i|\varphi _2\varphi _1|},$$ (9) where $`\gamma _i`$ is the switching rate of $`i`$-th EF in units of the Josephson frequency $`\omega _J=2eV/\mathrm{}`$. It is related to the dimensional switching rate $`\mathrm{\Gamma }`$ as $`\gamma \mathrm{}\mathrm{\Gamma }/2eV`$. Depending on the construction of the SQPC there can be any number of EFs which are “in range” to influence the transparency. In junctions which are very small it is probable that only one single EF will be in the vicinity of the contact. In this case, the coupling constant $`A`$ and the switching rate $`\mathrm{\Gamma }`$ can be directly evaluated from the measured telegraph noise intensity in the normal state, $$S(\tau )I(t+\tau )I(t)_tI(t)_t^2.$$ Indeed, the current through a single mode QPC at low temperatures can be expressed according to the Landauer formula as $`I(t)=2e^2VD(t)/h`$. Consequently, the random telegraph noise intensity is equal to $$S(\tau )=(2e^2VA/h)^2\mathrm{exp}(2\mathrm{\Gamma }\tau ),$$ (10) and both $`A`$ and $`\mathrm{\Gamma }`$ can be extracted from measured $`S(\tau )`$. A possible approach for extracting these model parameters from noise measurements in the case of many fluctutators in the QPC area will be discussed later. ## III Single EF In the case of a very small contact it is possible to consider only one EF and put $`i=1`$. We start by decomposing Eq. (1) as $`E(\varphi )=E_+(\varphi )+E_{}(\varphi )\xi (\varphi ),`$ (11) $`E_\pm {\displaystyle \frac{1}{2}}\left[E(\varphi |D_1)\pm E(\varphi |D_1)\right],`$ (12) where $`D_{\pm 1}`$ are the two different values the transparency fluctuates between. Further, we assume that both $`\gamma `$ and $`g_{fs}E_{}/eV`$ are much smaller than the reduced inter-level distance $`(E_+\mathrm{\Delta })/eV`$. This means that all deviations in time are much longer than the Andreev level formation time, which is of the order $`\mathrm{}/2\mathrm{\Delta }`$. Fortunately, the expression above, Eq. (12), is linear in $`\xi `$ and we can write the effect of the EF as an additive contribution to the accumulated phase, Eq. (8), without making any approximations. Namely, $`\mathrm{\Phi }\mathrm{\Phi }+\mathrm{\Phi }_f`$, with $$\mathrm{\Phi }_f=\frac{1}{2}_{\varphi _A}^{\varphi _B}g_{fs}(\varphi )\xi (\varphi )d\varphi .$$ (13) with $`g_{fs}(\varphi )=E_{}(\varphi )/eV`$. After averaging over the realizations of the random process $`\xi (t)`$, the expression (7) for the MAQI current is replaced by $$I_{\mathrm{dc}}=I_0[1W\mathrm{cos}(2\mathrm{\Theta }+2\mathrm{\Phi })],$$ (14) where $`We^{2i\mathrm{\Phi }_f}`$ contains the dephasing. Without dephasing $`W=1`$ and the effect of the phase $`\mathrm{\Phi }(\varphi _A,\varphi _B)`$ is a modulation of the dc current between 0 and $`2I_0`$, the modulation depth $$(I_{\mathrm{max}}I_{\mathrm{min}})/(I_{\mathrm{max}}+I_{\mathrm{min}})=|W|.$$ (15) being equal to one. This modulation of the current is the interference effect utilized in the MAQI. When dephasing enters, the quantity $`W(\varphi _A,\varphi _B)`$ will decrease and the modulation $`|W|`$ of the current envelope, i. e. the interference pattern, will in turn decrease. To facilitate the calculation of the dephasing term, $`W`$, we define the auxiliary function, $$\mathrm{\Psi }(\varphi )=\mathrm{exp}\left[i_{\varphi _A}^\varphi d\varphi ^{}g_{fs}(\varphi ^{})\xi (\varphi ^{})\right].$$ (16) The quantity of interest, $`W`$, is related to $`\mathrm{\Psi }(\varphi )`$ as $`W\mathrm{\Psi }(\varphi _B)`$. Further, it can be shown, cf. with Ref. , that Eq. (16) satisfies the differential equation $$\frac{\mathrm{d}^2\mathrm{\Psi }}{\mathrm{d}\varphi ^2}+\left[2\gamma \frac{\mathrm{d}\mathrm{ln}g_{fs}(\varphi )}{\mathrm{d}\varphi }\right]\frac{\mathrm{d}\mathrm{\Psi }}{\mathrm{d}\varphi }+g_{fs}^2(\varphi )\mathrm{\Psi }=0,$$ (17) with the initial conditions $$\mathrm{\Psi }(\varphi _A)=1,\mathrm{d}\mathrm{\Psi }/\mathrm{d}\varphi |_{\varphi =\varphi _A}=0.$$ (18) Let us consider the following two cases in more detail: (i) the “slow EF”, $`\gamma g_{fs}`$, which corresponds to low temperatures, and (ii) the “fast EF”, $`\gamma g_{fs}`$, which corresponds to relatively high temperatures. ### A Slow switching EF In the low-temperature limit the EF will slowly switch between its two states, and we can let $`\gamma 0`$ in Eq. (17). Further, by introducing the function, $$\mathrm{\Xi }(\varphi )=_{\varphi _A}^\varphi d\varphi g_{fs}(\varphi )\mathrm{\Psi }(\varphi ),$$ (19) and applying the initial conditions (18) we obtain the integral equation, $$\mathrm{\Xi }^2+\mathrm{\Psi }^2=1.$$ The solution follows as $`\mathrm{\Psi }^{(0)}(\varphi )=\mathrm{cos}\mathrm{\Phi }(\varphi )`$ where $`\mathrm{\Phi }(\varphi )\left(_{\varphi _A}^\varphi d\varphi ^{}g_{fs}(\varphi ^{})\right)`$. Consequently, $`W^{(0)}=\mathrm{cos}\mathrm{\Phi }_\zeta `$ with $$\mathrm{\Phi }_\zeta \mathrm{\Phi }(\varphi _B)=_{\varphi _A}^{\varphi _B}d\varphi g_{fs}(\varphi ).$$ (20) A better approximation can be found by looking for the solution in the form $`\mathrm{\Psi }^{(1)}(\varphi )=u(\varphi )\mathrm{\Psi }^{(0)}(\varphi )`$ and assuming $`u(\varphi )`$ to be a slow function. Neglecting the second derivative of $`u(\varphi )`$ we obtain the differential equation $$\mathrm{d}u/\mathrm{d}\varphi =\gamma \eta (\varphi )u$$ (21) for $`u(\varphi )`$ with $$\eta (\varphi )\left(1+\frac{1}{2g_{fs}^2\mathrm{tan}\mathrm{\Phi }(\varphi )}\frac{\mathrm{d}g_{fs}}{\mathrm{d}\varphi }\right)^1.$$ (22) Defining $$\upsilon _\zeta \frac{1}{2\zeta }_{\varphi _A}^{\varphi _B}\eta (\varphi )d\varphi $$ (23) we arrive at the following expression for the oscillating part of the current, $`W\mathrm{cos}(\mathrm{\Phi })=e^{2\gamma \upsilon _\zeta \zeta }\mathrm{cos}(\mathrm{\Phi }_\zeta )\mathrm{cos}(\mathrm{\Phi })`$ (24) $`=e^{2\gamma \upsilon _\zeta \zeta }\left[\mathrm{cos}(\mathrm{\Phi }+\mathrm{\Phi }_\zeta )+\mathrm{cos}(\mathrm{\Phi }\mathrm{\Phi }_\zeta )\right]/2.`$ (25) For a constant $`g_{fs}`$, $`\upsilon _\zeta =1`$ and $`\mathrm{\Phi }_\zeta =2g_{fs}\zeta `$, where $`\zeta =(\varphi _B\varphi _A)/2`$ is equal to half the distance between the resonance positions, see Fig. 1. In the general case these quantities are increasing functions of $`\zeta `$. At $`\gamma =0`$, the current is split into two interference patterns of equal magnitude shifted by the phase $`\mathrm{\Phi }_\zeta `$. A plot of $`\mathrm{\Phi }_\zeta `$ as a function of $`\zeta `$ is presented in Fig. 2. The transparency $`D`$ and the strength of the fluctuator are shown in the inset. There is no dephasing, only a distortion of the interference pattern, the modulation being $`|W|=|\mathrm{cos}\mathrm{\Phi }_\zeta |`$. This splitting into two patterns of equal magnitude follows from the assumption that the occupation probability is the same for the two EF states. The general case of arbitrary probabilities for the EF states can be solved numerically. At finite $`\gamma `$ dephasing takes place and the amplitude of the interference oscillations decreases by $`e^{2\gamma \upsilon _\zeta \zeta }`$. The physical reason of dephasing is the finite life time of an EF in a given state. We have calculated $`W`$ for $`\gamma =0.1`$, comparing the analytical approximation above against a numerical solution of Eq. (17), cf. with Fig. 3. (The plot for the analytical case when $`D=0.6,A=0.05`$ is missing because of numerical difficulties). ### B Fast switching EF In the case of fast switching the differential equation (17) can be approximated as, $$\frac{\mathrm{d}\mathrm{\Psi }}{\mathrm{d}\varphi }=\frac{g_{fs}^2(\varphi )}{2\gamma }\mathrm{\Psi }(\varphi ).$$ (26) The solution in this case is easily obtained, as $$W=\mathrm{\Psi }(\varphi _B)=\mathrm{exp}[K(\varphi _B,\varphi _A)],$$ (27) with $$K(\varphi _B,\varphi _A)=\frac{1}{2\gamma }_{\varphi _A}^{\varphi _B}d\varphi g_{fs}^2(\varphi ).$$ (28) Here we find an exponential decay of the interference term. Expression (28) describes an effect which is similar to motional narrowing of spectral lines. When the EF fluctuates rapidly enough comparied to the “energy resolution” $`E_{}/\mathrm{}`$, influence from the difference between two EF states is smeared and dephasing will be of a diffusive character, with en effective, time-dependent, diffusion constant $`g_{fs}^2/2\gamma `$. A calculation of the factor $`K`$ for $`\gamma =3`$ is shown in Fig. 4 for three different sets of values of the junction transparency and the EF’s strength. The effect of dephasing from flicker noise will decrease for higher bias voltages, since a higher bias, through the Josephson relation, leads to a shorter time between the resonances. To get an idea about the general case of arbitrary switching rates, it is instructive to consider a $`\varphi `$-independent $`g_{fs}`$. This approximation is only valid when $`|\mathrm{d}E(\varphi )/\mathrm{d}\varphi |`$ is small, but should provide the general behavior of $`W`$. In this limit Eq. (17) can be solved analytically, resulting in $`W=e^{2\gamma \zeta }[\mathrm{cosh}\left(2\zeta \sqrt{\gamma ^2g_{fs}^2}\right)+`$ (29) $`{\displaystyle \frac{\gamma }{\sqrt{\gamma ^2g_{fs}^2}}}\mathrm{sinh}\left(2\zeta \sqrt{\gamma ^2g_{fs}^2}\right)].`$ (30) From this expression one can conclude that there will always enter an exponential decay of the interference current, except for the limit of $`\gamma 0`$. In this limit there should only be an additional oscillation added to the current as a function of $`\zeta `$ since for $`\gamma |g_{fs}|`$, the hyperbolic functions behave as trigonometric functions. ## IV Many EF’s Let us consider a large number of EFs with varying switching rates distributed in the contact area. For simplicity, we shall assume that only the fluctuators with inter-level spacings $`U_ikT`$ are important, and that their distribution is uniform, $`𝒫_U(U)=P_0𝒱`$. Here $`𝒱`$ is the sample volume. Further, we assume that the switching rates $`\gamma _i`$ are the same for both transition directions (up och down) between the EF’s levels. This assumption is natural because the ratio between the corresponding transition rates is $`\mathrm{exp}(U_i/kT)`$. Within the assumptions discussed above, the final results are substantially simplified while preserving the essential depedence on temperature and the resonance position. These approximations agree with a general theory developed in Ref. for the case of dephasing by two-level systems (TLS) in glasses. The first step now is to linearize the SQPC’s transparency with respect to $`\xi _i`$ as $`DD+D_f(t)`$. This allows us to once again find an additive contribution to the accumulated phase, Eq. (8), which in this case will be, $$\mathrm{\Phi }_f\underset{i}{}A_i_{\varphi _A}^{\varphi _B}\xi _i(\varphi )g_{fm}(\varphi )d\varphi .$$ (31) Here we have defined $`g_{fm}(\varphi )=(1/2eV)\mathrm{d}E(\varphi )/\mathrm{d}D`$. In the same manner as in Sect. III, but this time averaging over $`A`$’s, $`\gamma `$’s and $`\xi ^{}s`$, we can express the modified MAQI current through expression (14) with $$W=e^{i_iA_i_{\varphi _A}^{\varphi _B}g_{fm}(\varphi )\xi _i(\varphi )d\varphi }_{A,\gamma ,\xi }.$$ (32) To approximate this average we use the Holtsmark method which is valid in the limit of many fluctuators, $`N=P_0𝒱kT1`$. This allows us to rewrite Eq. (32) as the average over the contributions from single EFs, $$W_s(A,\gamma )=\mathrm{exp}\left(iA_{\varphi _A}^{\varphi _B}g_{fm}(\varphi )\xi (\varphi )d\varphi \right)_\xi $$ (33) as $$W\mathrm{exp}\left(P_0𝒱kT1W_s(A,\gamma )_{A,\gamma }\right).$$ (34) Since the number of EFs is assumed to be large, to keep dephasing at a reasonable level it is important to keep $`1W_s`$ small. With the solutions for $`W_s`$ found in Sect. III the average $`1W_s`$ remains to be calculated. To calculate this average one has to specify the distributions of the parameters $`A`$ and $`\gamma `$. The simplest and most natural assumption is that these two quantities are not correlated. Consequently, the distribution $`𝒫(A,\gamma )`$ can be decoupled as $`𝒫_A(A)𝒫_\gamma (\gamma )`$. To specify the distribution $`𝒫_A`$ let us assume that the EFs are uniformly distributed in space. An EF behaves like a dipole, either electric or elastic, this allows us to specify it‘s interaction strength as $`A(r)=A_0/r^3`$, where $`r`$ is the distance between the contact and a given EF, while $`A_0`$ is a coupling constant dependent on a specific interaction mechanism. Note that the quantity $`A_0`$ has dimension of volume. Within this model we arrive at the normalized distribution function $`𝒫_A(A)=4\pi A_0/3𝒱A^2`$ (see Appendix B for details). The distribution $`𝒫_\gamma (\gamma )`$ is specified in a manner which is commonly used in glasses. Namely, the logarithm of $`\gamma `$ is assumed to be uniformly distributed. Hence, $`𝒫_\gamma (\gamma )\gamma ^1`$, see Appendix B. To normalize it let us take into account that for a given energy spacing $`U`$ there is a maximal switching rate. Since we are interested in the fluctuators with $`U_ikT`$, we can specify the maximal switching rate as $`\gamma _T`$, which is a function of the temperature. The actual temperature dependence is determined by the specific interaction mechanism between the EF and its environment. If the transitions between the EF states are caused by interaction with phonons, then $`\gamma _TT^3`$ while if the transitions are caused by the electrons excitations, then $`\gamma _TT`$ Therefore , the normalized distribution can be specified as $`𝒫_\gamma (\gamma )=(\gamma )^1`$, where $`=\mathrm{ln}(\gamma _T/\gamma _{\mathrm{min}})1`$. Here we have introduced the minimal switching rate, $`\gamma _{\mathrm{min}}`$. To express the decay in a more clear form let us introduce the dimensionless frequency $`\nu _d`$ corresponding to the interaction strength for an EF separated from the contact by an average distance to the active fluctuators, $`\overline{r}(4\pi P_0kT/3)^{1/3}`$, divided by the Josephson energy $`2eV`$. We can specify $`\nu _d`$ as $$\nu _d=4\pi P_0kTA_0/3=A_0/\overline{r}^3.$$ (35) The decay rate $`𝒦=\mathrm{ln}W`$ is then given by the expression $$𝒦=\frac{\nu _d}{}_0^{\mathrm{}}\frac{\mathrm{d}A}{A^2}_{\gamma _{\mathrm{min}}}^{\gamma _T}\frac{\mathrm{d}\gamma }{\gamma }\left[1W_s(A,\gamma )\right].$$ (36) Note that $`g_{fs}`$ has to be replaced with $`Ag_{fm}`$ in the expressions for $`W_s`$ in the many EF case, because of differences in notation. To estimate the amount of dephasing let us take into account that the asymptotic expressions for $`W_s`$, found in (25) and (27) for slow and fast switching respectively, match at $$\gamma \stackrel{~}{\upsilon }_\zeta (A\overline{g}_\zeta )^2\eta _\zeta /2\gamma ,$$ (37) where $`\stackrel{~}{\upsilon }_\zeta `$ differs from $`\upsilon _\zeta `$ defined in Eq. (23) by the replacement $`g_{fs}Ag_{fm}`$ in the differential equation (21), while $$\overline{g}_\zeta =\frac{1}{2\zeta }_{\varphi _A}^{\varphi _B}g_{fm}(\varphi ),\eta _\zeta \frac{1}{2\zeta \overline{g}_\zeta ^2}_{\varphi _A}^{\varphi _B}g_{fm}^2(\varphi )d\varphi .$$ (38) Note that $`\stackrel{~}{\upsilon }_\zeta `$ is a function of $`A`$, $`\stackrel{~}{\upsilon }_\zeta (A)`$. Consequently, Eq. (37) should be treated as an *equation* to determine the characteristic value of the coupling $`A`$. Defining the solution of Eq. (37) as $`A_\zeta `$ and splitting the integration over $`A`$ in Eq. (36) as $$_{\gamma _{\mathrm{min}}}^{\gamma _T}\frac{\mathrm{d}\gamma }{\gamma }\left(_0^{A_\zeta }+_{A_\zeta }^{\mathrm{}}\right)\frac{\mathrm{d}A}{A^2}\left[1W_s(A,\gamma )\right]$$ (39) we can use the expression (27) in the first interval and the expression (25) in the second one. Since both integrals are determined by $`A_\zeta `$ we arrive at the estimate $$\mathrm{ln}W3\nu _d\zeta f_\zeta ,$$ (40) where $`f_\zeta \overline{g}_\zeta \sqrt{\eta _\zeta \stackrel{~}{\upsilon }_\zeta (A_\zeta )}`$ is some function of $`\zeta `$, rather smooth if $`D`$ is not close to 1. We see that the interference pattern decays exponentially with an increasing distance, $`2\zeta `$, between the resonances. Generally the $`\zeta `$-dependence of $`f_\zeta `$ can be calculated numerically for a given transparency $`D`$ using the analytical expressions (23) and (37). For a constant $`g_{fm}`$, $`f_\zeta =g_{fm}`$. We do not analyze here the function $`f_\zeta `$ in detail. ### A Non-optimal EFs In the previous sections it has been assumed that the system size is infinite. A consequence of this assumption is that, independent of temperature, the EFs which have the strongest effect on the junction transparency will always be included in the estimates. From the method used to obtain the general estimate for many EFs, Eq. (40), one can conclude that the EFs which have the most effective coupling fulfill $`AA_\zeta `$, this corresponds to a spatial distance $`r_\gamma =(A_0/A_\zeta )^{1/3}(A_0\overline{g}_\zeta /\gamma )^{1/3}`$. A further point is that the rate $`\gamma `$ is confined to the interval between $`\gamma _{\mathrm{min}}`$ and $`\gamma _T`$. Thus we have actually assumed that the size of the region where EFs reside has a size greater than $`r_{\mathrm{max}}(A_0\overline{g}_\zeta /\gamma _{\mathrm{min}})^{1/3}`$, and that there is no “excluded region” without EFs near the contact with the size less that $`r_{\mathrm{min}}(A_0\overline{g}_\zeta /\gamma _T)^{1/3}`$. What happens if this “optimum” EF is out of the range? This can occur if the system is limited in size, or if there is a specifically pure region around the contact. #### 1 Role of finite size of the sample Let us first discuss the role of a finite size, $`R`$ , of the region containing EFs. If $`Rr_{min}=(A_0/A_\zeta )^{1/3}`$ then all EFs will act as “slow” ones. To estimate $`\mathrm{ln}W`$ in this case one can use the expression $`W_s=\mathrm{cos}\mathrm{\Phi }_\zeta `$ for the whole integration region over $`A`$ in Eq. (36), arriving at $$\mathrm{ln}W=2\nu _d\overline{g}_\zeta \zeta F(2A_0\overline{g}_\zeta \zeta /R^3).$$ (41) Here $$F(z)=_z^{\mathrm{}}\frac{1\mathrm{cos}x}{x^2}dx,$$ (42) which is a decreasing function of its argument. This expression (41) is valid if its right-hand side is less the right-hand side of Eq. (40). At intermediate values of $`R`$, $$(A_0\overline{g}_\zeta /\gamma _{min})^{1/3}R(A_0\overline{g}_\zeta /\gamma _T)^{1/3},$$ as in the previous case, only the second integral in the expression (39) does exist, lowest limit should also be replaced by $`A_0\overline{g}_\zeta /R^3`$. However, the approximation $`W_s=\mathrm{cos}\mathrm{\Phi }_\zeta `$ is not valid any more. Using the approximation (25) one can obtain $$W=\mathrm{exp}\left[\nu _d\zeta f_\zeta \frac{\mathrm{ln}(\gamma _TR^3/A_0\overline{g}_\zeta )}{\mathrm{ln}(\gamma _T/\gamma _{\mathrm{min}})}\right].$$ (43) This means that when the size of system is limited the amount of dephasing can be less than estimated for an infinite system. #### 2 Role of the spacer Let us now discuss the role of a “pure” region (spacer) near the SQPC where there are no EFs. If a typical diameter $`r_0`$ of such region is large enough, such that $`r_0r_{\mathrm{max}}=(A_0\overline{g}_\zeta /\gamma _{\mathrm{min}})^{1/3}`$, then all EFs act as “fast” ones. The single EF solution, $`W_s`$, in this limit is given by Eq. (27) in Sect. III B. After calculating the average in Eq. (34) we arrive at $$\mathrm{ln}W(\nu _d\overline{g}_\zeta /)\sqrt{4\pi \zeta \eta _\zeta /\gamma _{\mathrm{min}}}.$$ (44) It is difficult to estimate the actual amount of dephasing. We have to restrict our conclusions to interpreting how the amount of dephasing will change depending on the parameters $`\gamma _T`$, $`\gamma _{\mathrm{min}}`$, $`\zeta `$ and $`T`$. We already know that dephasing will increase with temperature. Generally one can also say that dephasing will increase with $`\zeta `$. However, at large enough temperatures when $`\gamma _{\mathrm{min}}`$ appears large enough, the dephasing will decrease. The free parameter of the theory, $`A_0`$, can be estimated only roughly through comparison with the noise measurements in the normal state. To map the parameter $`A_0`$ to the noise let us employ the theory of flicker noise in a QPC to the case of a single mode contact. According to that theory, results for the noise intensity $`S(\tau )`$ are substantially dependent on the relationship between the maximal and minimal distances between the EFs and the QPC. The simplest case, which is quite realistic, is when these distances are of the same order of magnitude. When $`\mathrm{\Gamma }_T^1|\tau |\mathrm{\Gamma }_{\mathrm{min}}^1`$ the noise intensity can be expressed as, cf. with Ref. , $$S(\tau )\left(\frac{2e^2V}{h}\right)^2\left(\frac{4\pi P_0kTA_0}{3}\right)^2\frac{\mathrm{ln}(1/\mathrm{\Gamma }_{\mathrm{min}}|\tau |)}{\mathrm{ln}(\mathrm{\Gamma }_T/\mathrm{\Gamma }_{\mathrm{min}})}.$$ (45) By obtaining estimates for $`\mathrm{\Gamma }_{T/\mathrm{min}}`$ from noise spectra in the normal state one can, in principle, estimate the coupling parameter $`A_0`$. A key point is to make measurements of both the MAQI interference pattern and the normal-state noise spectra in a rather large frequency range. This combination does not look too simple. ## V Conclusions We have presented a method for investigating the influence of flicker noise in the junction transparency of a SQPC on coherent Andreev states. This is done by estimating the effect of these fluctuations on the so-called microwave-activated quantum interferometer (MAQI) . For a small contact when only a single EF is in range to affect the junction transparency there can be either a distortion or a decay of the MAQI interference pattern. A distortion appears for very slow switching rates of the EF and a weak decay (dephasing) for fast rates. It is possible to confirm our model in the fast switching limit experimentally. The only unknown parameter is the switching rate $`\gamma `$ which can be found by measuring the telegraph noise of the contact in the normal state. This is best done by driving the system into the normal state with a magnetic field, since $`\gamma `$ is temperature dependent. When $`\gamma `$ is known it is then possible to calculate the amount of dephasing and compare with experimental results. In the limit when the influence of many EFs has to be considered we have arrived at more general results. It is not possible to make any exact predictions since the distribution and coupling strength of the EFs are sample dependent. However, our calculations show that in the presence of many EFs there will always be an exponential decay of the MAQI interference pattern. The strength of this dephasing should be about the same for all switching rates. One exception is when there is an impurity-free region near the SQPC, in this case dephasing will decrease for higher rates. Finally, we note that this paper together with work in Ref. presents a framework which can be used to investigate the coupling of a SQPC to its electromagnetic environment. ## Acknowledgements We are grateful to M. Jonson for fruitful discussions. Support from the SSF program “Quantum Devices and Nanostructures” and CAS, the Centre for Advanced Study in Oslo, Norway is acknowledged. ## A Derivation of equation for $`W`$ In this appendix the differential equation, Eq. 17, for $`W`$ is derived. The following is valid for the random telegraph process $`\xi (t)`$, $$\xi (\varphi _1)\xi (\varphi _2)\mathrm{}\xi (\varphi _n)=e^{2\gamma |\varphi _2\varphi _1|}\xi (\varphi _3)\xi (\varphi _4)\mathrm{}\xi (\varphi _n),$$ (A1) when $`\varphi _1\varphi _2\mathrm{}\varphi _n`$. Further, all averages with an odd number of $`\xi `$:s are equal to zero. This follows from the assumption of equal probability for both EF states. To facilitate the calculation of $`W`$, the auxiliary function $$\mathrm{\Psi }(\varphi )=\mathrm{exp}\left[i_{\varphi _A}^\varphi d\varphi ^{}g_{fs}(\varphi ^{})\xi (\varphi ^{})\right]$$ (A2) is defined. By expanding the auxiliary function in a Taylor series and utilizing the relation in Eq. A1 above, it is possible to rewrite the auxiliary function as $`\mathrm{\Psi }(\varphi )=`$ (A3) $`1+i^2{\displaystyle _{\varphi _A}^\varphi }d\varphi _1{\displaystyle _{\varphi _A}^{\varphi _1}}d\varphi _2g_{fs}(\varphi _1)g_{fs}(\varphi _2)e^{2\gamma |\varphi _2\varphi _1|}\mathrm{\Psi }(\varphi _2).`$ (A4) A derivation with respect to $`\varphi `$ leads to the following integro-differential equation $$\frac{\mathrm{d}\mathrm{\Psi }}{\mathrm{d}\varphi }=g_{fs}(\varphi )_{\varphi _A}^\varphi d\varphi ^{}g_{fs}(\varphi ^{})e^{2\gamma |\varphi ^{}\varphi |}\mathrm{\Psi }(\varphi ^{}),$$ (A5) and a second derivation provides the following differential equation $$\frac{\mathrm{d}^2\mathrm{\Psi }}{\mathrm{d}\varphi ^2}+\left[2\gamma \frac{\mathrm{d}\mathrm{ln}g_{fs}(\varphi )}{\mathrm{d}\varphi }\right]\frac{\mathrm{d}\mathrm{\Psi }}{\mathrm{d}\varphi }+g_{fs}^2(\varphi )\mathrm{\Psi }=0,$$ (A6) with the initial conditions $$\mathrm{\Psi }(\varphi _A)=1,\mathrm{d}\mathrm{\Psi }/\mathrm{d}\varphi |_{\varphi =\varphi _A}=0.$$ (A7) This equation makes it possible to find approximate analytical solutions for $`W`$ in the single EF case. ## B Distributions, $`𝒫_\gamma (\gamma )`$ and $`𝒫_A(A)`$ Here we outline the derivation for the distribution functions $`𝒫(\gamma )`$ and $`𝒫(A)`$ which are necessary when averaging in the many EF case. Let us begin with $`𝒫(A)`$, where $`A`$ is the coupling strength between an EF and the contact. By assuming that an EF can be modeled with a dipole field, we can state that $`AA_0/r^3`$ where $`r`$ is the distance from the EF. The normalized distribution is then found through the following integral, $`𝒫(A)=𝒱^1d𝐫\delta (AA_0/r^3)`$ as $$P(A)=(4\pi /3𝒱)(A_0/A^2).$$ (B1) To find the distribution of $`\gamma `$ a closer look at the structure of the two level system (TLS) which is assumed for the EFs is necessary. The Hamiltonian for a TLS can be $`H=\mathrm{\Delta }\sigma _z\mathrm{\Lambda }\sigma _x`$, where $`\mathrm{\Delta }`$ is the energy difference between the minima of the two states and $`\mathrm{\Lambda }`$ is the tunneling coupling between the two states. After diagonalization the excitation energy is found to be $`U=\sqrt{\mathrm{\Delta }^2+\mathrm{\Lambda }^2}`$. By assuming that the tunneling coupling decays exponentially we have that $`𝒫(\mathrm{\Lambda })1/\mathrm{\Lambda }`$. Further, the hopping rate $`\gamma `$ depends on $`(\mathrm{\Lambda }/U)^2`$ and a term $`U^3/\mathrm{}E_c^2`$ where the latter term comes from assuming that hopping is phonon mediated ($`E_c`$ is the parameter characterizing the coupling energy between the phonons and the EFs). Within these assumptions $`𝒫(\gamma )`$ is given by, $`𝒫_\gamma (\gamma )`$ $`=`$ $`{\displaystyle d\mathrm{\Delta }d\mathrm{\Lambda }P(\mathrm{\Lambda })\delta (U^2\mathrm{\Delta }^2\mathrm{\Lambda }^2)}`$ (B3) $`\times \delta (\gamma {\displaystyle \frac{\mathrm{\Lambda }^2}{U^2}}{\displaystyle \frac{U^3}{\mathrm{}E_c^2}}){\displaystyle \frac{1}{\gamma }}.`$ After normalization we have, $$P(\gamma )=\frac{1}{\gamma },$$ (B4) where $`=\mathrm{ln}(\gamma _T/\gamma _{\mathrm{min}})`$ with $`\gamma _{T,\mathrm{min}}`$ being limiting values for the EF switching rate $`\gamma `$. In the general case $`𝒫=𝒫(U,\gamma ,A)`$ and the number of EFs will be given by $`N=dUd\gamma dA𝒫(U,\gamma ,A)`$. By assuming a constant distribution for $`U`$, which we label $`𝒫_0`$, we arrive at the number of EFs as $`N=P_0𝒱kT`$. The final distribution function for $`A`$ and $`\gamma `$ is then $$𝒫(A,\gamma )=\frac{4\pi P_0kT}{3}\frac{A_0}{A^2}\frac{1}{\gamma }.$$ (B5)
warning/0003/cs0003008.html
ar5iv
text
# Consistency Management of Normal Logic Program by Top-down Abductive Proof Procedure ## Introduction Knowledge base is always subject to change since an environment around the knowledge base is not guaranteed to be stable forever and even some error might be included at the initial stage. Therefore, study of revision of knowledge base is very important(???????). (?) and (?) consider a revision of monotonic theories and there are a lot of researches in this direction (see (?) for a survey). (?) and (?) consider an update of nonmonotonic theories to derive a given goal or a given observation. (?) and (?) consider a revision of nonmonotonic theories which is more related to a revision of monotonic theories studies (??); they consider a revision when inconsistency arises at addition of rules <sup>1</sup><sup>1</sup>1(?) relate the latter approach with the former approach by introducing “anti-explanation of contradiction.” In this paper, we follow the latter approach. Revision of nonmonotonic theories is especially important for AI, since it is very rare that commonsense reasoning can be represented as a monotonic theory. However, revision of nonmonotonic theories is more complicated than revision of monotonic theory. In monotonic theory, if some addition of knowledge or observation leads to inconsistency, then we can avoid inconsistency by deleting a part of knowledge base. On the other hand, we might add a piece of assumptions since deletion leads to inconsistency. Consider the following program. $`runs(X)car(X),broken(X).`$ $`car(c_1).`$ $`car(c_2).`$ $``$ means “negation as failure”. The first rule says that if $`X`$ is a car and $`X`$ is not known to be broken, $`X`$ should run. Since there is no information about $`broken(c_1)`$ and $`broken(c_2)`$ in the current program, $`runs(c_1)`$ and $`runs(c_2)`$ are derived. Suppose, however, that we add a rule “$`runs(c_1)`$” meaning that car $`a`$ does not run <sup>2</sup><sup>2</sup>2$``$ means contradiction.. Then, we have no stable model, that is, we are in an inconsistent situation. To fix this inconsistency, we have (at least) two possible ways. 1. We simply discard the default rule of car $`a`$: $$runs(c_1)car(c_1),broken(c_1).$$ 2. We derive $`broken(c_1)`$ since if we assumed $`broken(c_1)`$, then contradiction would occur and thus, we have a reason to assume $`broken(c_1)`$. The first revision is contraction widely used in belief revision of monotonic theories (?) but the second is special for nonmonotonic theories such as normal logic programs. In monotonic theories, addition of formula can not help to restore consistency, but in nonmonotonic theories addition can help. This phenomena were firstly observed in Doyle’s justification TMS and he introduced dependency-directed backtracking. Moreover, in monotonic theories, contraction or deletion of formula can not produce any inconsistency, whereas in nonmonotonic theories, deletion can cause inconsistency. Therefore, we need more functions for revision in nonmonotonic theories than monotonic ones. In this paper, we propose a top-down procedure using abduction to compute revision for a normal logic program when there exists no stable model. Our idea of using abduction for revision is as follows. We introduce two kinds of abducibles one of which represents a deletion of each retractable rule and the other of which represents an addition of each addable rule. For a retractable rule, we add negation of a corresponding abducible in the body of the rule so that if an instance of abducible is assumed then an instance of rule corresponding with the instance of abducible is no longer applicable. For an addable rule, we add a corresponding abducible in the body of the rule so that if an instance of abducible is assumed then an instance of rule corresponding with the instance of abducible becomes applicable. Then, in order to compute such abducibles to specify revision, we show that we can use a modification of Satoh and Iwayama’s query evaluation procedure on stable models (?) which is a combination of integrity constraint checking (?) and abductive procedure (?). In stead of starting with a subprocedure which show a derivation of positive literals, we start with a subprocedure for rule consistency checking to derive abducibles to specify revision. This procedure traverses rules of a program which is related with addition or deletion of a rule and we guarantee that a minimal revision can be found by selecting a minimal set of abducibles among all the sets of abducibles computed by the rule consistency checking procedure. ## Revision of Normal Logic Program Firstly, we define a revision framework as follows. In this paper, we consider a function-free normal logic program. We use domain closure axiom and unique name axiom so that constants in the language for a program are finite and denote distinct objects. We can easily extend our results to function-free extended logic programs by translating an extended logic program into a normal logic program proposed by (?). ###### Definition 1 A rule $`R`$ is of the form: $$HP_1,\mathrm{},P_j,N_1,\mathrm{},N_h$$ where $`H`$, $`P_1,\mathrm{},P_j,N_1,\mathrm{},N_h`$ are atoms. We call $`H`$ the head of the rule $`R`$ denoted as $`head(R)`$ and $`P_1,\mathrm{},P_j,N_1,\mathrm{},N_h`$ the body of the rule denoted as $`body(R)`$. If $`H=`$, we sometimes call the rule an integrity constraint. Let $`T`$ and $`T_{bck}`$ be sets of rules. A revision framework $``$ is a pair, $`T,T_{bck}`$ where $`T`$ can be divided into two sets of rules $`T_{pst}`$ and $`T_{tmp}`$. We call $`T_{pst}`$ a persistent part of $``$ and $`T_{tmp}`$ a temporal part of $``$ and $`T_{bck}`$ a backup part of $``$. $`T`$ expresses the current logic program which consists of $`T_{pst}`$ and $`T_{tmp}`$. $`T_{pst}`$ is an unchanged part which should always be satisfied such as integrity constraints whereas any part of $`T_{tmp}`$ can be retracted and any part of $`T_{bck}`$ can be added to restore consistency. Usage of $`T_{bck}`$ is inspired by back-up semantics proposed by (?). We use stable model semantics for the above program. ###### Definition 2 Let $`P`$ be sets of rules. We denote a set of ground rules obtained by replacing all the variables in every rule of $`P`$ by every element in the language as $`\mathrm{\Pi }_P`$. ###### Definition 3 Let $`M`$ be a set of ground atoms and $`\mathrm{\Pi }_P^M`$ be the following program. $`\mathrm{\Pi }_P^M=`$ $`\{HB_1,\mathrm{},B_l|`$ $`\mathrm{`}\mathrm{`}HB_1,\mathrm{},B_l,A_1,\mathrm{},A_h.^{\prime \prime }\mathrm{\Pi }_P`$ and $`A_iM`$ for each $`i=1,\mathrm{},h.\}`$ Let $`min(\mathrm{\Pi }_P^M)`$ be the least model of $`\mathrm{\Pi }_P^M`$. A stable model for a logic program $`P`$ is $`M`$ iff $`M=min(\mathrm{\Pi }_P^M)`$ and $`M`$. We say that $`P`$ is consistent if $`P`$ has a stable model. Now, we define a revised program in a revision framework. ###### Definition 4 Let $``$ be a revision framework, $`(T_{pst}T_{tmp}),T_{bck}`$. Let $`R_{new}`$ be a rule. Then, a revised program w.r.t. $``$ and $`R_{new}`$ is $`(T_{pst}\{R_{new}\})(\mathrm{\Pi }_{T_{tmp}}O)I`$ such that * $`O\mathrm{\Pi }_{T_{tmp}}`$ * $`I\mathrm{\Pi }_{T_{bck}}`$ * $`(T_{pst}\{R_{new}\})(\mathrm{\Pi }_{T_{tmp}}O)I`$ is consistent. We say for such $`O`$ and $`I`$ that a pair $`O,I`$ accomplishes revision of $`R_{new}`$ to $``$. A minimally revised program w.r.t. $``$ and $`R_{new}`$ is $`(T_{pst}\{R_{new}\})(\mathrm{\Pi }_{T_{tmp}}O)I`$ such that there is no revised program $`(T_{pst}\{R_{new}\})(\mathrm{\Pi }_{T_{tmp}}O^{})I^{}`$ such that $`I^{}I`$ and $`O^{}O`$ where $``$ is a strict inclusion. ###### Example 1 Let $`T_{pst}`$ be $`\{c(c_1).c(c_2).\}`$ and $`T_{tmp}`$ be $`\{r(X)c(X),b(X).\}`$ and $`T_{bck}`$ be $`\{b(X)c(X),r(X).\}`$ and $`R_{new}`$ be “$`r(c_1)`$”. Then, $`\mathrm{\Pi }_{T_{tmp},(T_{pst}\{R_{new}\})T_{tmp}T_{bck}}=`$ $`\{r(c_1)c(c_1),b(c_1).r(c_2)c(c_2),b(c_2)\}`$, and $`\mathrm{\Pi }_{T_{bck},(T_{pst}\{R_{new}\})T_{tmp}T_{bck}}=`$ $`\{b(c_1)c(c_1),r(c_1).b(c_2)c(c_2),r(c_2)\}`$. For the above revision, we have the following two minimally revised programs: 1. a program accomplished by $`(O_1,I_1)`$ where $`O_1=\{r(c_1)c(c_1),b(c_1).\}`$ and $`I_1=\{\}`$: $`(T_{pst}\{R_{new}\})`$ and $`r(c_2)c(c_2),b(c_2).`$ 2. a program accomplished by $`(O_2,I_2)`$ where $`O_2=\{\}`$ and $`I_2=\{b(c_1)c(c_1),r(c_1).\}:`$ $`(T_{pst}\{R_{new}\})`$ and $`T_{tmp}`$ and $`b(c_1)c(c_1),r(c_1).`$ There are other non-minimally revised programs, for example, one accomplished by $`O_1,I_2`$ or by $`O_1\{r(c_2)c(c_2),b(c_2).\},I_2\{b(c_2)c(c_2),r(c_2).\}`$. In the above example, we follow Giordano’s approach (?) where contrapositives of default rules are in the back-up part, but we can actually assume any rules which we think are appropriate for back-up rules when inconsistency occurs. To compute a revised program, we use a translation from a specification to an abductive logic program and compute a consistent generalized stable model for the translated program which denotes deletion and addition of rules. ###### Definition 5 (?). An abductive framework is a pair $`P,A`$ where $`A`$ is a set of predicate symbols, called abducible predicates and $`P`$ is a set of rules each of whose head is not in $`A`$. We call a ground atom for a predicate in $`A`$ an abducible. The semantics of abductive framework is based on a generalized stable model (?). The following is a definition of a generalized stable model which can manipulate abducibles in abductive logic programming. ###### Definition 6 Let $`P,A`$ be an abductive framework and $`\mathrm{\Theta }`$ be a set of abducibles. A generalized stable model $`M(\mathrm{\Theta })`$ is a stable model of $`P\{H|H\mathrm{\Theta }\}`$. We say that a model $`M(\mathrm{\Theta })`$ is a generalized stable model with a minimal set of abducibles $`\mathrm{\Theta }`$ if there is no generalized stable model $`M(\mathrm{\Theta }^{})`$ such that $`\mathrm{\Theta }^{}`$ is a proper subset of $`\mathrm{\Theta }`$. Now, we define a translation of a revision framework into an abductive framework as follows. ###### Definition 7 Let $``$ be a revision framework $`(T_{pst}T_{tmp}),T_{bck}`$. We firstly give a name to every rule in $`T_{tmp}`$ and $`T_{bck}`$ such as $$\varphi :HP_1,\mathrm{},P_j,N_1,\mathrm{},N_h.$$ where $`\varphi `$ is a name for the rule. A translation for a consistency management of $``$ (denoted as $`\tau ()`$) is a set of the following translation from $``$ to an abductive framework $`P,A`$ where * $`A=\{\varphi _{}^{}{}_{}{}^{}|\varphi `$ is a name of a rule in $`T_{tmp}\}`$ $`\{\varphi _{}^{+}{}_{}{}^{}|\varphi `$ is a name of a rule in $`T_{bck}\}`$ * We add every rule in $`T_{pst}`$ into $`P`$. * We translate every rule in $`T_{tmp}`$ with a name $`\varphi `$ $$\varphi :HP_1,\mathrm{},P_j,N_1,\mathrm{},N_h$$ into the following rule in $`P`$: $$HP_1,\mathrm{},P_j,N_1,\mathrm{},N_h,\varphi _{}^{}{}_{}{}^{}(𝐱)$$ where $`𝐱`$ is a tuple of variables in the clause. * We translate every rule in $`T_{bck}`$ with a name $`\varphi `$ $$\varphi :HP_1,\mathrm{},P_j,N_1,\mathrm{},N_h$$ into the following rule in $`P`$: $$HP_1,\mathrm{},P_j,\varphi _{}^{+}{}_{}{}^{}(𝐱),N_1,\mathrm{},N_h.$$ The following shows that an revised program corresponds with a generalized stable model. ###### Theorem 1 Let $``$ be a revision framework $`(T_{pst}T_{tmp}),T_{bck}`$ and $`R_{new}`$ be an added clause. $`(T_{pst}\{R_{new}\})(T_{tmp}T_{del}T_{new})`$ is a (minimally, resp.) revised program if and only if there is a generalized stable model of $`\tau ((T_{pst}\{R_{new}\})T_{tmp},T_{bck})`$ with a (minimal, resp.) set of abducibles $`\mathrm{\Theta }`$ s.t. * $`T_{del}=\{R|(\varphi _{}^{}{}_{}{}^{}(𝐱)\theta )\mathrm{\Theta }`$ where $`\varphi `$ is a name of $`RT_{tmp}\}`$. * $`T_{new}=`$ $`\{R\theta |(\varphi _{}^{+}{}_{}{}^{}(𝐱)\theta )\mathrm{\Theta }`$ where $`\varphi `$ is a name of $`RT_{bck}\}`$ $``$ $`\{head(R)body(R),(EQ(\theta _1)),\mathrm{},(EQ(\theta _n))|`$ $`(\varphi _{}^{}{}_{}{}^{}(𝐱)\theta _i)\mathrm{\Theta }`$ where $`\varphi `$ is a name of $`RT_{tmp}`$ and $`EQ(\theta _i)=((x_1=(x_1\theta _i))\mathrm{}(x_k=(x_k\theta _i)))`$ and $`𝐱=x_1,\mathrm{},x_k\}`$ We say that $`\mathrm{\Theta }`$ (minimally, resp.) realizes revision of $`R_{new}`$ to $``$. Note that in the above theorem, we delete whole rules related to inconsistency and then add modified rules with negation of conjunctions of disequality in the body of the deleted rules. The modified rules are logically equivalent to rules in Definition 4 since we assume domain closure axiom and unique name axiom. ###### Example 2 Consider the revision framework in Example 1. Let us give names to the rules in $`T_{tmp}`$ and $`T_{bck}`$ as follows: $`\varphi _1:r(X)c(X),b(X).`$ $`\varphi _2:b(X)c(X),r(X).`$ Then $`\tau (((T_{pst}\{R_{new}\})T_{tmp}),T_{bck})`$ is: * $`A=\{\varphi _{1}^{}{}_{}{}^{},\varphi _{2}^{+}{}_{}{}^{}\}`$. * $`P`$ becomes as follows: $`c(c_1).`$ $`c(c_2).`$ $`r(c_1)`$. $`r(X)c(X),b(X),\varphi _{1}^{}{}_{}{}^{}(X).`$ $`b(X)c(X),\varphi _{2}^{+}{}_{}{}^{}(X),r(X).`$ Then, we have two generalized models with minimal abducibles: 1. $`\mathrm{\Theta }=\{\varphi _{1}^{}{}_{}{}^{}(c_1)\}`$. Then, a minimally revised program is: $`(T_{pst}\{R_{new}\})`$ and $`r(X)Xa,c(X),b(X).`$ 2. $`\mathrm{\Theta }=\{\varphi _{2}^{+}{}_{}{}^{}(c_1)\}`$. Then, a minimally revised program is: $`(T_{pst}\{R_{new}\})`$ and $`T_{tmp}`$ and $`b(c_1)c(c_1),r(c_1).`$ ## Computing Revision by Abduction To compute a revision, it is sufficient to compute a generalized stable model of $`\tau ((T_{pst}\{R_{new}\})T_{tmp},T_{bck})`$. But, if we concern a minimal revision, we need to compute all the generalized stable models and then compare sets of abducibles in these models to choose minimal sets of abducibles. For this purpose, it is desirable to restrict sets of abducibles to be compared as small as possible. This can be done if we compute only revision related to $`R_{new}`$. For example, suppose that some of temporary rules are not relevant to inconsistency of addition of $`R_{new}`$. If we naively compute all the generalized stable models, then we have to compare all the combination of in/out of abducibles for these irrelevant rules. In order to avoid this kind of redundancy, we modify Satoh and Iwayama’s query evaluation procedure on stable models (?). Basically, we change the order of application of subprocedures so that we can use the procedure for consistency checking. We impose rules in a revision framework must be range-restricted, that is, any variable in a rule $`R`$ must occur in $`pos(R)`$. However, any rule can be translated into range-restricted form by inserting a new predicate “$`dom`$” describing Herbrand universe for every non-range-restricted variable in the rule. Before showing our procedure to compute revision, we need the following definitions. Let $`l`$ be a literal. Then, $`\overline{l}`$ denotes the complement of $`l`$. ###### Definition 8 Let $`P`$ be a logic program. A set of resolvents w.r.t. a ground literal $`l`$ and $`T`$, $`resolve(l,P)`$ is the following set of rules: $`resolve(l,P)=`$ $`\{(L_1,\mathrm{},L_k)\theta |`$ $`l`$ is negative and $`HL_1,\mathrm{},L_kP`$ and $`\overline{l}=H\theta `$ by a ground substitution $`\theta \}`$ $`\{(HL_1,\mathrm{},L_{i1},L_{i+1},\mathrm{},L_k)\theta |`$ $`HL_1,\mathrm{},L_kP`$ and $`l=L_i\theta `$ by a ground substitution $`\theta \}`$ ###### Definition 9 Let $`P`$ be a logic program. A set of deleted rules w.r.t. a ground literal $`l`$ and $`P`$, $`del(l,P)`$, is the following set of rules: $`del(l,P)=\{(HL_1,\mathrm{},L_k)\theta |`$ $`HL_1,\mathrm{},L_kP`$ and $`\overline{l}=L_i\theta `$ by a ground substitution $`\theta \}`$ ###### Definition 10 Let $`P`$ be a logic program and $`P^{}`$ be an abducible-and-negation-removed program obtained by removing all integrity constraints in $`P`$ and all the negative literals and abducibles in the body of remaining rule and $`min(P^{})`$ be the least minimal model of $`P^{}`$.We define a relevant ground program $`\mathrm{\Omega }_P`$ for $`P`$ as follows: $`\mathrm{\Omega }_P=\{HB_1,\mathrm{},B_k,A_1,\mathrm{},A_m\mathrm{\Pi }_P|`$ $`B_imin(P^{})\text{ for each }i=1,\mathrm{},k.\}`$ We briefly explain our procedure. Our procedure consists of 4 subprocedures, $`rule\mathrm{\_}con(R,\mathrm{\Delta })`$, $`derive(p,\mathrm{\Delta })`$, $`literal\mathrm{\_}con(l,\mathrm{\Delta })`$, and $`deleted\mathrm{\_}con(R,\mathrm{\Delta })`$ where $`p`$ is a non-abducible atom and $`\mathrm{\Delta }`$ is a set of ground literals already assumed and $`l`$ is a ground literal and $`R`$ is a rule. $`rule\mathrm{\_}con(R,\mathrm{\Delta })`$, $`literal\mathrm{\_}con(l,\mathrm{\Delta })`$, and $`deleted\mathrm{\_}con(R,\mathrm{\Delta })`$ return union of $`\mathrm{\Delta }`$ and a set of ground literals which are assumed during the execution of the subprocedures. $`derive(p,\mathrm{\Delta })`$ return the above union and a substitution for $`p`$ which are made during the execution of $`derive(p,\mathrm{\Delta })`$. In the procedure, we have a select operation and a fail operation. The select operation expresses a nondeterministic choice among alternatives. The fail operation expresses immediate termination of an execution with failure. Therefore, a procedure succeeds when its inner calls of subprocedures do not encounter fail. We say a subprocedure succeeds with (a substitution $`\theta `$ and) a set of assumptions $`\mathrm{\Delta }`$ when the subprocedure successfully returns ($`\theta `$ and) $`\mathrm{\Delta }`$. Our procedure firstly starts from $`rule\mathrm{\_}con(R_{new},\{\})`$. $`rule\mathrm{\_}con(R_{new},\{\})`$ checks the consistency of a rule $`R_{new}`$ with a program $`P\{R_{new}\}`$. We can show the consistency of addition of $`R_{new}`$ by showing one of the following. 1. A literal $`l`$ in $`body(R_{new})`$ can be falsified. To do so, we invoke subprocedure $`literal\mathrm{\_}con`$ for $`\overline{l}`$. 2. Every positive literal $`p`$ in $`body(R_{new})`$ can be made true and every negative and every abducible literal $`l`$ can be consistently assumed and $`head(R_{new})`$ consistent. To do so, we invoke subprocedure $`derive`$ for $`p`$ and $`literal\mathrm{\_}con`$ for $`l`$ and $`head(R_{new})`$. The informal specification of the other 3 subprocedures is as follows. 1. $`literal\mathrm{\_}con(l,\mathrm{\Delta })`$ checks the consistency of a ground literal $`l`$ with $`P\{R_{new}\}`$ and $`\mathrm{\Delta }`$. To show the consistency for assuming $`l`$, we add $`l`$ to $`\mathrm{\Delta }`$; then, we check the consistency of resolvents and deleted rules w.r.t. $`l`$ and $`P\{R_{new}\}`$. 2. $`derive(p,\mathrm{\Delta })`$ searches a rule $`R`$ of $`p`$ in a program $`P\{R_{new}\}`$ whose body can be made true with a ground substitution $`\theta `$ under a set of assumptions $`\mathrm{\Delta }`$. To show that every literal in the body can be made true, we call $`derive`$ for non-abducible positive literals in the body. Then, we check the consistency of other literals in the body with $`P\{R_{new}\}`$ and $`\mathrm{\Delta }`$. Note that because of the range-restrictedness, other literals in $`R`$ become ground after all the calls of $`derive`$ for non-abducible positive literals. 3. $`deleted\mathrm{\_}con(R,\mathrm{\Delta })`$ checks if a deletion of $`R`$ does not cause any contradictions with $`P\{R_{new}\}`$ and $`\mathrm{\Delta }`$. To show the consistency of the implicit deletion of $`R`$, it is sufficient to prove that the head of every ground instance $`R\theta `$ in $`\mathrm{\Omega }_{P\{R_{new}\}}`$ can be made either true or false. Now, we describe a complete specification of the subprocedures in Figure 1 and Figure 2. In Figures, we denote a set of non-abducible positive literals, non-abducible negative literals, and abducibles (either negative or positive) in a rule $`R`$ as $`pos(R)`$, $`neg(R)`$ and $`abd(R)`$ respectively, and we denotes empty substitution as $`\epsilon `$, and $`\theta _i\sigma _i`$ expresses a composition of two substitutions $`\theta _i`$ and $`\sigma _i`$. The following theorem on correctness for rule checking can be derived from correctness on query evaluation procedure of (?). ###### Theorem 2 Let $`P,A`$ be a consistent abductive framework. Suppose $`rule\mathrm{\_}con(R_{new},\{\})`$ succeeds for $`P`$ with $`\mathrm{\Delta }`$, then there is a generalized stable model $`M(\mathrm{\Theta })`$ for $`PR_{new},A`$ such that $`\mathrm{\Theta }`$ includes positive abducibles in $`\mathrm{\Delta }`$. par The above theorem only guarantees that $`R`$ is consistent with $`P`$ and the procedure produces some abducibles included in a generalized stable model. To compute revision, however, we must have the stronger result that $`\mathrm{\Delta }`$ includes all the necessary $`\varphi _{}^{+}{}_{}{}^{}`$’s and $`\varphi _{}^{}{}_{}{}^{}`$’s. Actually, we can guarantee this by the following theorem. ###### Theorem 3 Let $``$ be a revision framework $`(T_{pst}T_{tmp}),T_{bck}`$ such that $`T_{pst}T_{tmp}`$ is consistent and $`R_{new}`$ be an added rule. Suppose $`rule\mathrm{\_}con(R_{new},\{\})`$ succeeds for $`\tau ()`$ with $`\mathrm{\Delta }`$, then, a set of positive abducibles in $`\mathrm{\Delta }`$ realizes revision of $`R_{new}`$ to $``$. The following theorem means that if we can search exhaustively in selecting the rules or cases and there is a generalized stable model whose abducibles minimally realizes a revision for addition of rule $`R_{new}`$, then we can find such a set of abducibles by our procedure. Note that this property is always guaranteed if a program is a finite propositional program or has finite constant symbols and no function symbols. ###### Theorem 4 Let $``$ be a revision framework $`(T_{pst}T_{tmp}),T_{bck}`$ and $`R_{new}`$ be an added rule. Suppose that every selection of rules or cases terminates for $`rule\mathrm{\_}con(R_{new},\{\})`$ with either success or failure for $`\tau ()`$. If $`\mathrm{\Theta }`$ minimally realizes revision of $`R_{new}`$ to $`T`$, then there is a selection of rules and cases such that $`rule\mathrm{\_}con(R_{new},\{\})`$ succeeds with $`\mathrm{\Delta }`$ where a set of abducibles in $`\mathrm{\Delta }`$ is equivalent to $`\mathrm{\Theta }`$. Note that we cannot guarantee that positive abducibles of every $`\mathrm{\Delta }`$ always corresponds with a minimal revision. This problem is inherited from Satoh’s procedure in that it is not guaranteed for the procedure to produce a minimal abducibles. However, using the procedure, we can restrict sets of abducibles related with inconsistency and, thus, considered sets of abducibles to choosing minimal sets are smaller than sets of abducibles from a naive calculation of all the generalized stable models. ###### Example 3 Consider the revision framework in Example 1. The following are two sequences of main calls of subprocedures for $`rule\mathrm{\_}con(R_{new},\{\})`$ to $`\tau (U)`$ shown in Example 2. In the following, $`rc`$, $`lc`$, $`dr`$ and $`dc`$ corresponds with $`rule\mathrm{\_}con`$, $`literal\mathrm{\_}con`$, $`derive`$ and $`deleted\mathrm{\_}con`$ respectively, and indexes in the front express a nesting structure of the calls. Note that existence of “$`c(c_2)`$.” in $`T_{pst}`$ does not influence these derivations. Sequence 1 (for $`I_1,O_1`$ in Example 1) $`rc((r(c_1)),\{\})`$ 1 $`lc(r(c_1),\{\})`$ 1.1 $`rc((c(c_1),b(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),\{r(c_1)\})`$ 1.1.1 $`dr(c(c_1)\{r(c_1)\})`$ select $`c(c_1).`$ 1.1.1.2 $`lc(c(c_1)\{r(c_1)\})`$ 1.1.1.2.1 $`rc((r(c_1)b(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),\{c(c_1),r(c_1)\})`$ 1.1.1.2.1.1 $`lc(b(c_1),\{c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.1 $`rc((r(c_1)c(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),`$ $`\{c(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.1.1.1.1 $`dr(\varphi _{1}^{}{}_{}{}^{}(c_1),\{c(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.1.1.1.1.1 $`lc(\varphi _{1}^{}{}_{}{}^{}(c_1),\{c(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.1.1.1.1.1.1 $`dc((r(c_1)c(c_1),b(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),`$ $`\{\varphi _{1}^{}{}_{}{}^{}(c_1),c(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.1.1.2 $`rc((c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),r(c_1)),`$ $`\{\varphi _{1}^{}{}_{}{}^{}(c_1),c(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.1.1.2.1 $`lc(\varphi _{2}^{+}{}_{}{}^{}(c_1),`$ $`\{\varphi _{1}^{}{}_{}{}^{}(c_1),c(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.1.1.2.1.1 $`dc((b(c_1)c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),r(c_1)),`$ $`\{\varphi _{1}^{}{}_{}{}^{}(c_1),c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),b(c_1),r(c_1)\})`$ 1.1.1.2.2 $`rc((b(c_1)\varphi _{2}^{+}{}_{}{}^{}(c_1),r(c_1)),`$ $`\{\varphi _{1}^{}{}_{}{}^{}(c_1),c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),b(c_1),r(c_1)\})`$ 1.2 $`rc((b(c_1)c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1)),`$ $`\{\varphi _{1}^{}{}_{}{}^{}(c_1),c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),b(c_1),r(c_1)\})`$ 1.3 $`dc((r(c_1)),`$ $`\{\underset{¯}{\varphi _{1}^{}{}_{}{}^{}(c_1)},c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),b(c_1),r(c_1)\})`$ Sequence 2 (for $`I_2,O_2`$ in Example 1) $`rc((r(c_1)),\{\})`$ 1 $`lc(r(c_1),\{\})`$ 1.1 $`rc((c(c_1),b(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),\{r(c_1)\})`$ 1.1.1 $`dr(c(c_1)\{r(c_1)\})`$ select $`c(c_1).`$ 1.1.1.2 $`lc(c(c_1)\{r(c_1)\})`$ 1.1.1.2.1 $`rc((r(c_1)b(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),\{c(c_1),r(c_1)\})`$ 1.1.1.2.1.1 $`dr(b(c_1),\{c(c_1),r(c_1)\})`$ select $`b(c_1)c(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),r(c_1).`$ 1.1.1.2.1.1.2 $`dr(c(c_1),\{c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.3 $`lc(\varphi _{2}^{+}{}_{}{}^{}(c_1),\{c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.3.1 $`rc((b(c_1)c(c_1),r(c_1)),`$ $`\{\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.3.1.1 $`lc(b(c_1),\{\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.3.1.1.1 $`dc((r(c_1)c(c_1),b(c_1),\varphi _{1}^{}{}_{}{}^{}(c_1)),`$ $`\{b(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.4 $`lc(r(c_1),\{b(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.1.1.2.1.1.5 $`lc(b(c_1),\{b(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.1.1.2.2 $`rc((b(c_1)\varphi _{2}^{+}{}_{}{}^{}(c_1),r(c_1)),`$ $`\{b(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.2 $`rc((b(c_1)c(c_1),`$ $`\varphi _{2}^{+}{}_{}{}^{}(c_1)),\{b(c_1),\varphi _{2}^{+}{}_{}{}^{}(c_1),c(c_1),r(c_1)\})`$ 1.3 $`dc((r(c_1)),\{b(c_1),\underset{¯}{\varphi _{2}^{+}{}_{}{}^{}(c_1)},c(c_1),r(c_1)\})`$ ## Related Work There are works on calculation method of updates (??). (?) propose a top-down procedure to compute view updates in a database for proving a given goal, but it is not applicable to updating a normal logic program in general. (?) give a bottom-up procedure for an update of an acyclic program to explain given observations in the perfect model of an acyclic program and therefore, cannot apply to a normal logic program which has multiple stable models. Moreover, using bottom-up computation would lead to an irrelevant derivation to an added rule. In (?), translation from an update framework (?) to an extended logic program is provided <sup>3</sup><sup>3</sup>3Recently, (?) independently proposes exactly the same technique as our translation to show correspondence their extended abduction and ordinary abduction.. Differences between our translation and their translation are as follows. * They give a translation to compute an update to explain a goal whereas we consider a revision to avoid inconsistency of addition of a rule. * They introduce a new predicate symbol in stead of abducibles. This makes their translation rather complex. If we translate our translated abductive logic program to a new normal logic program by a method proposed in (?), the new normal logic program would be the same as their program. * They consider addition/deletion of the whole rules to derive a given observation. That is, instead of considering deletion/addition of parts of $`\mathrm{\Pi }_{T_{tmp}}`$/$`\mathrm{\Pi }_{T_{bck}}`$ in Definition 4, they propose deletion/addition of parts of $`T_{tmp}`$/$`T_{bck}`$ . At least, however, to handle exception of integrity constraints in software engineering (?), we believe that our fine-grained approach is better since we would like to keep consistent part of integrity constraints for further checking of other data when some instances cause inconsistency. See the detailed discussion in (?). There are many procedures to compute stable models, generalized stable models or abduction. If we use a bottom-up procedure for our translated abductive logic program to compute all the generalized stable models naively, then sets of abducibles to be compared would be larger since abducibles of irrelevant temporary rules and addable rules with inconsistency will be considered. Therefore, it is better to compute abducibles related with inconsistency. To our knowledge, top-down procedure which can be used for this purpose is only Satoh and Iwayama’s procedure since we need a bottom-up consistency checking of addition/deletion of literals during computing abducibles for revision. This task is similar to integrity constraint checking in (?) and Satoh and Iwayama’s procedure includes this task. ## Conclusion In this paper, we propose an abductive top-down procedure to compute a minimal revised program which traverses only relevant parts of the program to the added rule. It is done by translating a revision framework of a normal logic program into an abductive logic program. As a future work, we would like to find an efficient method of computing a minimal revision directly by combining our top-down procedure and ATMS-like method of memorizing justifications of revisions. ## Acknowledgments This research is partly supported by Grant-in-Aid for Scientific Research on Priority Areas, “Principles for Constructing Evolutionary Software”, The Ministry of Education, Japan. We also thank the anonymous referees for valuable comments on this paper.
warning/0003/cond-mat0003052.html
ar5iv
text
# Vortex-glass phases in type-II superconductors11footnote 1Accepted for publication in Advances in Physics ## 1 Introduction Since its discovery by Kammerlingh Onnes in 1911, superconductivity has attracted generations of physicists. In 1935 Fritz and Heinz London (see e.g. \[London 1950\] (?)) developed a very successful phenomenological theory which describes both the perfect conductivity as well as the perfect diamagnetism of superconductors. As discussed later by \[London 1950\] (?) this theory can be motivated by considering superconductivity as a phenomenon characterized by long-range order of momentum $`𝐩`$. Bohr-Sommerfeld quantization $`𝐩_s𝑑𝐬=nh`$ on a torus gives fluxoid quantization \[London 1950\]. \[Ginzburg and Landau 1950\] (?) combined London’s electrodynamics of a superconductor with Landau’s theory of phase transitions, creating a powerful phenomenological description of superconductivity. The transition to the superconducting phase corresponds here to the breaking of the U(1) symmetry of the complex order parameter $`\mathrm{\Psi }`$ and the appearance of off-diagonal long-range order (ODLRO). In a pioneering work \[Abrikosov 1957\] (?) showed the existence of a second type of superconductors which (for sufficiently strong external magnetic field) allows for a penetration of quantized magnetic flux in the form of vortex lines, which form a triangular lattice, reducing the perfect diamagnetism and creating a source for dissipation due to the motion of the vortex-line core driven by the transport current. In this case the continuous translational symmetry of the system is broken in addition to the $`U(1)`$ symmetry. Effects from thermal fluctuations, although studied already since 1960, were considered to be extremely small because of the large correlation length and low transition temperatures of conventional superconductors \[Ginzburg 1961\]. To keep the superconducting properties, vortices have to be prevented from moving by pinning centers. An early theory of pinning of isolated vortex lines \[Anderson and Kim 1964\] shows the absence of dissipation only at zero temperature. Thermally activated hopping leads to a small but finite dissipation at low temperatures (as compared to the height of energy barriers). \[Larkin 1970\] (?) extended this theory to the Abrikosov vortex-line lattice, showing the destruction of its translational long-range order. Although – in principle – the Abrikosov phase could thence be considered not to really differ from a pinned vortex-line liquid (and hence from the normal phase), the generic phase diagram of conventional superconductors was assumed to practically be that of \[Abrikosov 1957\] (?) with a now finite correlation length of the vortex-line array. In the 1980s this picture was changed by two initially unrelated developments: the discovery of high-$`T_c`$ superconductors by \[Bednorz and Müller 1986\] (?) and the much better understanding of random system, in particular of spin glasses and of random-field systems (for a recent review see \[Young 1998\] (?)). In the high-$`T_c`$ superconductors with their elevated transition temperatures and pronounced anisotropy, fluctuation effects became now very important as can be seen for instance from the observed melting of the vortex-line lattice \[Cubitt et al. 1993, Zeldov et al. 1995\]. Moreover, for pure systems it was demonstrated \[Moore 1989, Moore 1992, Ikeda et al. 1992\] that thermal fluctuations – which prohibit true long-range translational order of the vortex-line lattice (VLL) only in $`d2`$ dimensions – destroy the ODLRO of the gauge invariant order parameter in the Abrikosov phase even in higher dimensions. Thus, in three dimensions thermal fluctuations restore the $`U(1)`$ symmetry of the Ginzburg-Landau Hamiltonian but nevertheless allow for the existence of a vortex lattice. This finding is paralleled by an earlier observation of \[Schafroth 1955\] (?) that an external magnetic field above a critical strength destroys Bose-Einstein condensation of an ideal Bose gas which still shows some remanent diamagnetic moment. For systems with disorder the idea emerged that despite of the destruction of true translational long-range order the system could show a phase with some kind of glassy long-range order, the “vortex glass” \[Fisher 1989\]. Because of the residual rigidity in the vortex-line array, arbitrarily large energy barriers now exist, leading to a highly non-linear resistivity \[Feigel’man and Vinokur 1990, Fisher 1989, Feigel’man et al. 1989, Nattermann 1990, Fisher et al. 1991a\] $$\rho (j)e^{(j_\mathrm{t}/j)^\mu }$$ (1.1) where $`\mu `$ denotes an exponent $`0\mu 1`$ and $`j_\mathrm{t}`$ a threshold current. Since the linear resistivity vanishes, the system is truly superconducting. In the following years vortex pinning and depinning as well as flux creep under the action of an external current was investigated by many researchers to a great extent. A brilliant summary of the results of these efforts till 1994 is given in the extensive review article by \[Blatter et al. 1994\] (?) (see also \[Brandt 1995\] (?), \[Gammel et al. 1998\] (?), \[Giamarchi and Le Doussal 1998\] (?)). It is not the intention of the present article to provide an updated version of these reviews by discussing the results obtained since then. Instead, we want to focus here mainly on one particular aspect of the theory, namely on the discussion of the equilibrium phase diagram of weakly disordered type-II superconductors in an external magnetic field. We want to demonstrate that the notion of a ‘vortex glass’ is not a blurred expression for the hopelessly intricate situation in a disordered system (as some physicists may still claim), but that it has a well defined meaning. The knowledge of the equilibrium properties is also important for the proper understanding of situations close to equilibrium, e.g. for the discussion of flux creep under the influence of a small external current. Despite of the conclusion common to all authors mentioned above of expecting a non-analytic current-density dependence of the resistivity in the glassy phase it seems to be indicated to refer here also to the differences between these approaches. \[Fisher 1989\] (?) and \[Fisher et al. 1991a\] (?) in their definition of the glassy phase started from correlation functions measuring ODLRO and focused on the long-range (glassy) order of phases (see chapter 2.4). On the other hand, \[Feigel’man et al. 1989\] (?), \[Nattermann 1990\] (?) and subsequently \[Korshunov 1993\] (?) and \[Giamarchi and Le Doussal 1994\] (?) considered primarily the glassy order of the vortex-line array, i.e., they focused on the positions of the vortex lines. In both cases, the expression ‘vortex glass’ was used. The above mentioned differences in the lower critical dimensions for the breaking of the $`U(1)`$ and of the translational symmetry in pure systems however suggests, that there may be also different lower critical dimensions for a phase-coherent and a positional vortex glass. In this review we concentrate mainly on the positional glass. If, in particular, the positional vortex glass is free of large dislocation loops such that an elastic description of the vortex array is possible, the positional vortex glass will be called elastic vortex glass. The most prominent example of an elastic vortex glass is its three-dimensional version: the so-called ‘Bragg-glass’ (for a recent brief review see \[Giamarchi and Le Doussal 1998\] (?)). Whether non-elastic vortex-glass phases exist is still unclear. Systems with columnar disorder which leads to the formation of the so-called ‘Bose-glass’ phases (see e.g. \[Täuber and Nelson 1997\] (?)) will be completely neglected in this review since their physics is substantially different from that for systems with point disorder, to which we restrict ourselves here. However, we also include the discussion of various phases found for driven systems far from equilibrium, which share some features with the equilibrium phase diagram. The article is organized as follows. In chapter 2 we present a brief summary of the Ginzburg-Landau theory of type-II superconductors and a short discussion of the influence of thermal fluctuations and of the effect of point disorder in the critical region. We also define the different types of vortex-glass order and give a brief account of results obtained for models with strong disorder – the so-called gauge glasses. In chapter 3 we review the behaviour of a single vortex line and its generalizations – $`D`$-dimensional directed manifolds – in a random potential. This simple, but not at all trivial system allows for a discussion of different aspects of glassiness of a system. Chapter 4 is devoted to the superconducting film in a parallel field, a geometry which allows for a very detailed description of the vortex glass phase as well as of the transition to the normal phase both for the static and dynamic quantities. In chapter 5 we discuss an impure superconducting film in a field perpendicular to the film plane. It turns out that dislocations destroy the positional vortex-glass phase in this geometry. The ‘Bragg-glass’ phase of a bulk superconductor as well as its stability with respect to dislocations is considered in chapter 6. A short account of recent activities on driven vortex lattices in impure superconductors is presented in chapter 7. We close the paper with a brief summary of the results of this article (chapter 8). The appendix contains some technicalities and a list of recurrent symbols. ## 2 Ginzburg-Landau description In this chapter we give a very brief introduction into the mean-field theory and the effects arising from thermal and disorder fluctuations in type-II superconductors in the framework of the Ginzburg-Landau theory. Since there is extensive (and partially contradicting) literature on thermal effects it is impossible to include all related references. However, we attempt to include the most recent articles on the subject which may serve as more comprehensive guides to further references. ### 2.1 The Ginzburg-Landau model In 1950 Ginzburg and Landau proposed a phenomenological description of superconductors by introducing a two-component order parameter $`\mathrm{\Psi }(𝐫)=|\mathrm{\Psi }(𝐫)|e^{i\varphi (𝐫)}`$ which couples in a gauge-invariant form to the magnetic field described by the vector potential $`𝐀(𝐫)`$ \[Ginzburg and Landau 1950\]. The density $`n_\mathrm{s}(𝐫)`$ of superconducting charge carriers (i.e., of the Cooper pairs), which is a central quantity of the earlier London theory \[London and London 1935\], is related to $`\mathrm{\Psi }(𝐫)`$ by $`n_\mathrm{s}(𝐫)=|\mathrm{\Psi }(𝐫)|^2`$. The Ginzburg-Landau (GL) free energy is given by $`_{\mathrm{GL}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle }d^dr\{\beta (|\mathrm{\Psi }|^2+{\displaystyle \frac{\alpha }{\beta }})^2`$ (2.1) $`+{\displaystyle \frac{\mathrm{}^2}{m}}|(i\mathbf{}{\displaystyle \frac{2\pi }{\mathrm{\Phi }_0}}𝐀)\mathrm{\Psi }|^2+{\displaystyle \frac{1}{4\pi }}(\mathbf{}𝐀𝐇)^2\},`$ where $`\mathrm{\Phi }_0=hc/2e`$ denotes the flux quantum, $`\alpha (T)(TT_{\mathrm{c0}})`$, $`T_{\mathrm{c0}}`$ is the mean-field transition temperature, $`𝐇`$ is the external field, and $`m`$ denotes the mass of a Cooper pair. The GL free energy is characterized by two basic length scales, the coherence length $`\xi `$ and the penetration depth $`\lambda `$, which are related to the parameters of $`_{\mathrm{GL}}`$ by $`\xi ^2(T)`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2}{2m|\alpha (T)|}},`$ (2.2a) $`\lambda ^2(T)`$ $`=`$ $`{\displaystyle \frac{mc^2}{4\pi |\mathrm{\Psi }_0|^2(2e)^2}}.`$ (2.2b) Here $`|\mathrm{\Psi }_0|^2=|\alpha |/\beta `$ denotes the saturation value of $`|\mathrm{\Psi }|^2`$ in a homogeneous current free state for $`T<T_{\mathrm{c0}}`$ and $`𝐇=\mathrm{𝟎}`$. For our further discussion it is convenient to use the following rescaling to introduce dimensionless quantities $`\mathrm{\Psi }^{}`$, $`𝐀^{}`$ and $`𝐫^{}`$ $`\mathrm{\Psi }`$ $`=`$ $`\mathrm{\Psi }^{}|\alpha /\beta |^{1/2},𝐀={\displaystyle \frac{\mathrm{\Phi }_0}{2\pi \xi }}𝐀^{},𝐫=𝐫^{}\xi .`$ (2.3) This leads to $`{\displaystyle \frac{1}{T}}_{\mathrm{GL}}`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}\left({\displaystyle \frac{|\tau _{\mathrm{c0}}|^{4d}}{2\mathrm{G}\mathrm{i}}}\right)^{1/2}{\displaystyle }d^dr^{}\{{\displaystyle \frac{1}{2}}(|\mathrm{\Psi }^{}|^2+\alpha /|\alpha |)^2`$ (2.4) $`+|(i\mathbf{}^{}𝐀^{})\mathrm{\Psi }^{}|^2+\kappa ^2(\mathbf{}^{}𝐀^{}𝐇/H_{\mathrm{c2}}^{\mathrm{MF}})^2\}.`$ Here we introduced the Ginzburg number $$\mathrm{Gi}\left(\frac{T}{4\pi \xi ^d(0)\beta |\mathrm{\Psi }_0(0)|^4}\right)^2=\left(\frac{T}{\xi ^d(0)H_\mathrm{c}^2(0)}\right)^2,$$ (2.5) the Ginzburg-Landau parameter $`\kappa \lambda /\xi `$ and the reduced temperature $`\tau _{\mathrm{c0}}(T_{\mathrm{c0}}T)/T_{\mathrm{c0}}`$. $`H_{\mathrm{c2}}^{\mathrm{MF}}\mathrm{\Phi }_0/2\pi \xi ^2`$ is the mean-field upper critical field and $`H_\mathrm{c}^2=4\pi \beta |\mathrm{\Psi }_0|^4`$ is the thermodynamic critical field. ### 2.2 Mean-field theory Within mean-field (MF) theory, the GL free energy has to be minimized with respect to the fields $`\mathrm{\Psi }(𝐫)`$ and $`𝐀(𝐫)`$. The resulting GL-equations $$(i\mathbf{}^{}𝐀^{})^2\mathrm{\Psi }^{}+\frac{\alpha }{|\alpha |}\mathrm{\Psi }^{}+|\mathrm{\Psi }^{}|^2\mathrm{\Psi }^{}=0$$ (2.6) and (for $`\alpha <0`$) $$\kappa ^2\frac{1}{|\mathrm{\Psi }^{}|^2}\mathbf{}(\mathbf{}𝐀^{})+𝐀^{}=\mathbf{}^{}\varphi $$ (2.7) then have to be solved with the appropriate boundary conditions. As is clear from equation (2.4), the only two parameters which will enter the solution in the bulk, are the GL-parameter $`\kappa `$, which plays the role of an inverse effective charge of the $`\mathrm{\Psi }`$ field, and the strength of the external magnetic field $`𝐇^{}=𝐇/H_{\mathrm{c2}}^{\mathrm{MF}}`$. For $`\kappa <1/\sqrt{2}`$ (type-I superconductors), mean-field theory yields for $`T<T_{\mathrm{c0}}`$ and $`H<H_\mathrm{c}(T)`$ a phase with vanishing resistance and perfect diamagnetism. The transition to the normal phase at $`H_\mathrm{c}(T)`$ is first order. For $`\kappa >1/\sqrt{2}`$ (type-II superconductors), on the other hand, perfect diamagnetism exists only up to the field $`H_{\mathrm{c1}}^{\mathrm{MF}}(H_{\mathrm{c2}}^{\mathrm{MF}}/2\kappa ^2)(\mathrm{ln}\kappa +0.08)`$. For larger fields, magnetic flux penetrates the sample in the form of quantized vortex lines, each carrying a flux quantum $`\mathrm{\Phi }_0`$. The energy per unit length of the vortex line is therefore given by $`\epsilon _\mathrm{l}=(\mathrm{\Phi }_0/4\pi )H_{\mathrm{c1}}\epsilon _0\mathrm{ln}\kappa `$ with $`\epsilon _0=(\mathrm{\Phi }_0/4\pi \lambda )^2`$ as the important energy scale per unit length. The vortex lines form a triangular ‘Abrikosov’ lattice \[Abrikosov 1957, Kleiner et al. 1964\] of spacing $`a_{\mathrm{}}=(2/\sqrt{3})^{1/2}a`$, where $`a(\mathrm{\Phi }_0/B)^{1/2}`$. The Abrikosov lattice (or ‘mixed’) phase shows both broken translational symmetry and off-diagonal long-range order (ODLRO), i.e., broken U(1) symmetry of the order parameter. Both broken symmetries vanish simultaneously if $`H`$ reaches $`H_{\mathrm{c2}}^{\mathrm{MF}}`$. One should however take into account that the correlation function for ODLRO (see chapter 2.3.2) $`\mathrm{\Psi }^{}(𝐫)\mathrm{\Psi }(𝐫^{})`$ – even if calculated in MF approximation – shows strong spatial variations due to the rapid change of the phase. Indeed, for a system of radius $`R`$ the tangential phase gradient on the boundary is of the order $`R/2a^2`$ which corresponds to a phase change of $`2\pi `$ on a distance $`l_\varphi 4\pi a^2/R`$ (with $`a100\mathrm{n}\mathrm{m}`$ and $`R1\mathrm{c}\mathrm{m}`$, $`l_\varphi 10^2\mathrm{nm}`$ which is smaller than an atom, see \[Brandt 1974\] (?)). Thus $`\mathrm{\Psi }^{}(𝐫)\mathrm{\Psi }(𝐫^{})`$ cannot be very meaningful as a physical observable. Quantities with a physical significance should be in particular gauge invariant. In the GL-description these are the amplitude $`|\mathrm{\Psi }|`$ of the order parameter, the ‘super-velocity’ $`\mathbf{}\varphi \frac{2\pi }{\mathrm{\Phi }_0}𝐀\mathbf{}\stackrel{~}{\varphi }`$ and the magnetic induction $`𝐁=\mathbf{}𝐀`$. All other physical quantities can be expressed in these fields, e.g. the current density can be written as $$𝐣=\frac{2e\mathrm{}}{m}|\mathrm{\Psi }|^2\mathbf{}\stackrel{~}{\varphi }.$$ (2.8) In treating the vortex system, two main approximations have been used: the lowest Landau level (LLL) approximation, which is valid sufficiently close to $`H_{\mathrm{c2}}^{\mathrm{MF}}`$ ($`H\frac{1}{3}H_{\mathrm{c2}}^{\mathrm{MF}}`$), and the London approximation, which is valid at intermediate and small fields $`B0.2H_{\mathrm{c2}}^{\mathrm{MF}}`$, where $`\xi a`$. The precise range of applicability of the LLL is still under debate (see e.g. \[O’Neill and Moore 1993\] (?), \[Li and Rosenstein 1999\] (?) and references therein). In the London approximation one neglects amplitude inhomogeneities, $`|\mathrm{\Psi }|=|\mathrm{\Psi }_0|`$, which leads to a diverging energy density at the vortex cores. The position $`𝐫_i(s)`$ of these cores is parameterized by the label $`i`$ of a vortex and the variable $`s`$ along the contour of the vortex lines. The Ginzburg-Landau Hamiltonian takes then the form $$_{\mathrm{London}}=\frac{1}{2}d^dr\left\{\frac{\mathrm{}^2}{m}|\mathrm{\Psi }_0|^2(\mathbf{}\varphi +\frac{2\pi }{\mathrm{\Phi }_0}𝐀)^2+\frac{1}{4\pi }(\mathbf{}𝐀𝐇)^2\right\}.$$ (2.9) This functional has to be regularized near the vortex cores, e.g. by excluding tubes of radius $`\xi `$ around the vortex cores from the volume integration. The phase $`\varphi (𝐫)`$ of the complex order parameter is a multivalued function since $`\varphi `$ changes by $`2\pi `$ along a path surrounding a vortex line. We decompose $`\varphi `$ now into a vortex part $`\varphi _\mathrm{v}`$ and a “spin-wave” part $`\varphi _{\mathrm{sw}}`$. The vortex part is assumed to fulfill the saddle point equation apart from the position of the vortices $`𝐫_i(s)`$. With the London gauge $`\mathbf{}𝐀=0`$ this yields $$\mathbf{}^2\varphi _\mathrm{v}(𝐫)=0,𝐫𝐫_i(s)$$ (2.10) and $$\mathbf{}(\mathbf{}\varphi _\mathrm{v})=2\pi 𝐦(𝐫).$$ (2.11) Here $`𝐦(𝐫)`$ denotes the vortex-density field $$𝐦(𝐫)=\underset{i}{}m_i𝑑s\frac{d𝐫_i(s)}{ds}\delta ^{(3)}(𝐫_i(s)𝐫),$$ (2.12) where the integration is along the vortex line $`i`$ which carries the vorticity $`m_i=\pm 1`$. If the spin-wave part $`\varphi _{\mathrm{sw}}(𝐫)`$ vanishes on the surface of the sample, $`\varphi _{\mathrm{sw}}`$ and $`\varphi _\mathrm{v}`$ decouple. Since the vector potential $`𝐀`$ appears only quadratically in $`H_{\mathrm{GL}}`$ it can be integrated out by using the saddle-point equation which is the second GL-equation in the phase-only approximation $$𝐀+\lambda ^2\mathbf{}(\mathbf{}𝐀)=\frac{\mathrm{\Phi }_0}{2\pi }\mathbf{}\varphi _\mathrm{v}.$$ (2.13) Taking the curl of (2.13) gives the modified London equation $$\lambda ^2\mathbf{}^2𝐁(𝐫)𝐁(𝐫)=\mathrm{\Phi }_0𝐦(𝐫),$$ (2.14) where $`𝐁=\mathbf{}𝐀`$ can now completely be expressed in terms of the vortex degrees of freedom given by the vortex density field $`𝐦(𝐫)`$, equation (2.12). The London Hamiltonian then takes the form $$_{\mathrm{London}}=\frac{1}{2}d^dr\left\{\frac{\mathrm{}^2}{m}|\mathrm{\Psi }_0|^2(\mathbf{}\varphi _{\mathrm{sw}})^2+\frac{\lambda ^2}{4\pi }(\mathbf{}𝐁)^2+\frac{1}{4\pi }(𝐁𝐇)^2\right\}.$$ (2.15) In most parts of this review we will use the London picture, since it remains valid in the vortex phases we will describe, in particular in the elastic glass phases. The elasticity theory of the Abrikosov lattice for an isotropic superconductor was worked out by Brandt (?, ?). The distortion of a vortex line is described by a two-component displacement field $`𝐮(𝐗,z)`$, where the lattice vector $`𝐗𝐗_{n,m}=((2n+m)a_{\mathrm{}}/2,m\sqrt{3}a_{\mathrm{}}/2)`$ denotes the rest position of the vortex line in the plane perpendicular to $`𝐇=H\widehat{𝐳}`$. In many cases one can go over to the continuum description: $`𝐮(𝐗,z)𝐮(𝐱,z)𝐮(𝐫)`$. On large scales $`L\lambda `$ the elastic energy of the vortex-line lattice is then given by $`_{\mathrm{el}}={\displaystyle \frac{1}{2}}{\displaystyle d^2xd^{d2}z\left\{c_{11}(\mathbf{}_{}𝐮)^2+c_{66}(\mathbf{}_{}𝐮)^2+c_{44}(\mathbf{}_{}𝐮)^2\right\}},`$ (2.16) where $`c_{11}c_{44}𝐁^2/4\pi `$. The Abrikosov phase is characterized in particular by a non-zero shear modulus $`c_{66}\frac{B\mathrm{\Phi }_0}{(4\pi \lambda )^2}\left(1\frac{B}{H_{\mathrm{c2}}^{\mathrm{MF}}}\right)^2`$, which vanishes both at $`H_{\mathrm{c2}}`$ and $`H_{\mathrm{c1}}`$, reaching a maximum of $`c_{66}H_{\mathrm{c1}}H_{\mathrm{c2}}/56\pi `$ in between. One should however take into account that the elastic free energy is in general non-local, which is expressed in a strong dispersion of $`c_{11}`$ and $`c_{44}`$ on scales smaller than $`\lambda `$ (see, e.g., \[Brandt 1991\] (?)). In the absence of pinning centres, the system in the Abrikosov phase behaves superconducting only for currents parallel to the magnetic field. For currents with components perpendicular to the field the Lorentz force drives the vortex-line array, which leads to metallic behaviour with resistivity $`\rho \rho _\mathrm{n}B/H_{\mathrm{c2}}`$ \[Bardeen and Stephen 1965\]. Here $`\rho _\mathrm{n}`$ is the resistivity of the normal phase. Before we come to the discussion of fluctuation effects, we want to consider a possible extension of the model (2.1) which describes an isotropic superconductor. High-$`T_\mathrm{c}`$ superconductors, however, are characterized by a pronounced layer structure, which results in an inhomogeneity and in a strong spatial anisotropy of the effective mass of the Cooper pair such that $`m`$ is replaced by $`M=m/ϵ^2`$ for electrons moving perpendicular to the layers. Typical values for $`ϵ`$ are $`ϵ_{\mathrm{YBCO}}0.16`$ and $`ϵ_{\mathrm{BSCCO}}10^2`$ for the two high-$`T_\mathrm{c}`$ materials YBCO and BSCCO. It was shown by \[Blatter et al. 1992\] (?) (for a more detailed discussion, see \[Blatter et al. 1994\] (?)), that in the case of $`\kappa 1`$ the result for a thermodynamic quantity $`𝒬(\vartheta ,H,T,\xi ,\lambda ,ϵ,\mathrm{\Delta })`$ of an anisotropic superconductor (where $`\vartheta `$ denotes the angle between the magnetic field direction and the $`xy`$ plane, and $`\mathrm{\Delta }`$ the strength of disorder) can be obtained from the corresponding result $`\stackrel{~}{𝒬}`$ for the isotropic system by the relation $`𝒬(\vartheta ,H,T,\xi ,\lambda ,ϵ,\mathrm{\Delta })=s_𝒬\stackrel{~}{𝒬}(ϵ_\vartheta H,T/ϵ,\xi ,\lambda ,\mathrm{\Delta }/ϵ),`$ (2.17) where $`ϵ_\vartheta ^2=ϵ^2\mathrm{cos}^2\vartheta +\mathrm{sin}^2\vartheta `$, $`s_V=s_E=s_T=ϵ`$ for volume, energy and temperature, and $`s_B=s_H=1/ϵ_\vartheta `$ for magnetic fields. Since $`T`$ and $`\mathrm{\Delta }`$ are increased by a factor $`1/ϵ`$ with respect to the isotropic system, it is clear that fluctuation effects, which will be considered in the following sections, are drastically enlarged (by a factor up to 100) in these materials. If the spatial anisotropy is so large that the coherence length $`\xi _z=\xi ϵ`$ in $`z`$ direction becomes of the order of the layer spacing $`s`$, the discreteness of the layer structure becomes relevant. In this case, new effects such as a decoupling of the layers \[for $`B>B_{2\mathrm{D}}(ϵ/s)^2\mathrm{ln}(s/\xi )`$\] may occur. The appropriate description is then the Lawrence-Doniach model \[Lawrence and Doniach 1971\]. We will not attempt to cover in this review also these particular features of strongly layered materials, instead we restrict ourselves in the following to the discussion of the isotropic superconductor, knowing that the results for the anisotropic case can be found from the relation (2.17). The neglect of layer-effects is also supported by the following argument: Our theoretical analysis will be mainly based on the elastic description of the vortex lattice and our main interest concerns features on large length scales. Sufficiently weak disorder will indeed effectively couple to the vortex lattice only on very large length scales, where the elasticity of the lattice may be described by local elasticity theory. Since within the London approach all information about anisotropy and even about the layered structure is encoded in the dispersion of the elastic constants, we expect that the large-scale properties are independent of these details (provided that one is in the parameter regime where the London approach is valid and that disorder is sufficiently weak). ### 2.3 Thermal fluctuations So far we have ignored the influence of fluctuations, i.e., of configurations which do not fulfill the GL equations. These can be taken into account if we interpret the GL free energy as an effective Hamiltonian from which the true free energy $`(T,𝐇)`$ has to be calculated as $`(T,𝐇)=T\mathrm{ln}\left({\displaystyle 𝒟\mathrm{\Psi }𝒟𝐀e^{_{\mathrm{GL}}/T}}\right).`$ (2.18) Since the only material independent common feature of type-II superconductors is flux quantization it is natural to build from $`\mathrm{\Phi }_0`$ and $`T`$ a characteristic length scale \[Fisher et al. 1991a\] $`\mathrm{\Lambda }_T{\displaystyle \frac{\mathrm{\Phi }_0^2}{16\pi ^2T}}210^8{\displaystyle \frac{\text{Å}}{T[\text{K}]}},`$ (2.19) which is the same for all materials. Since the energy per unit length of a vortex line (and hence also its stiffness constant $`\epsilon _{}`$, see below) is of the order $`\epsilon _0=(\mathrm{\Phi }_0/4\pi \lambda )^2`$, $`\mathrm{\Lambda }_T`$ denotes the length scale on which the mean squared displacement of a vortex line is of order $`\lambda `$. Since $`\mathrm{\Lambda }_T`$ is so large, thermal fluctuation effects are expected to be small (however, see our remark about strongly anisotropic systems in the previous chapter 2.2). In $`d=3`$ dimensions the Ginzburg number can be expressed as $`\mathrm{Gi}(\kappa \lambda (0)/\mathrm{\Lambda }_T)^2`$. As can be seen directly from (2.4) and (2.18) the contribution from fluctuations in $`\mathrm{\Psi }(𝐫)`$ and $`𝐀(𝐫)`$ will indeed be small, if both $`\mathrm{Gi}|\tau _{\mathrm{c0}}|^{4d}`$ and $`\kappa 1`$. #### 2.3.1 Zero external field In zero external field, $`𝐇=\mathrm{𝟎}`$, to begin with, it was shown by \[Halperin et al. 1974\] (?) that for type-I superconductors, where $`\kappa <1/\sqrt{2}`$ and typically $`\mathrm{Gi}1`$ (e.g., $`\mathrm{Gi}10^{13}`$ for aluminium), fluctuations in the vector potential $`𝐀(𝐫)`$ render the transition first order. For type-II superconductors, on the other hand, the situation is less clear. It has been argued that the transition remains second order \[Helfrich and Müller 1980, Dasgupta and Halperin 1981\]. In the high-$`T_\mathrm{c}`$ compounds with large values for $`\kappa `$ ($`\kappa _{\mathrm{YBCO}}100`$, $`\kappa _{\mathrm{BSCCO}}60`$) and large Ginzburg numbers ($`\mathrm{Gi}_{\mathrm{YBCO}}10^2`$, $`\mathrm{Gi}_{\mathrm{BSCCO}}1`$) fluctuations in the vector potential are weak compared to those of the order parameter. Then there exist two critical regions. In the outer critical region $`|\tau _\mathrm{c}|\mathrm{Gi}^{1/(4d)},|\tau _\mathrm{c}|(\mathrm{Gi}/\kappa ^4)^{1/(4d)},`$ (2.20) where $`\tau _\mathrm{c}(T_\mathrm{c}T)/T_\mathrm{c}`$ now denotes the reduced temperature with respect to the true transition temperature $`T_\mathrm{c}`$, fluctuations of the order parameter lead to an $`XY`$-like critical behaviour. Fluctuations in the vector potential can be neglected in this regime. Since the coherence length $`\xi (T)\xi (0)|\tau _\mathrm{c}|^\nu `$ with $`\nu =\nu _{XY}\frac{2}{3}`$ in $`d=3`$ dimensions increases now more strongly than the penetration depth $`\lambda \lambda _0|\tau _\mathrm{c}|^{\beta +\eta \nu /2}`$, the effective value of $`\kappa \xi ^{(4d)/2}`$ decreases until both lengths are of the same order. This signals a cross-over to a second critical regime with (probably) inverted $`XY`$ behaviour \[Helfrich and Müller 1980, Dasgupta and Halperin 1981\]. In this asymptotic regime $`\lambda `$ and $`\xi `$ scale in the same way with the correlation exponent $`\nu _{XY}`$ \[Olsson and Teitel 1999\]. It should be mentioned, however, that other scenarios have been proposed (for a recent discussion of earlier results see \[Kiometzis et al. 1995\] (?), \[Radzihovsky 1995a\] (?), \[Herbut and Tešanović 1996\] (?), \[Herbut 1997\] (?), \[Folk and Holovatch 1999\] (?), \[Nguyen and Sudbø 1999\] (?)). In $`d=2`$ dimensions fluctuations prevent the formation of a long-range ordered phase. As was shown by Pearl (?, ?), the effective London penetration depth $`L_\mathrm{s}=2\lambda ^2/s`$ diverges with decreasing $`s`$, where $`s`$ denotes the film thickness. Therefore fluctuations in the vector potential can be neglected and the system in zero external field shows a Kosterlitz-Thouless transition to a quasi-long-range ordered phase \[Doniach and Hubermann 1979, Halperin and Nelson 1979\]. #### 2.3.2 Finite external field Next we consider the case of finite external field. The most obvious effect of thermal fluctuations on the vortex-line lattice is melting \[Eilenberger 1967, Nelson 1988\]. Melting has been seen experimentally in YBCO \[Safar et al. 1992, Kwok et al. 1992, Charalambous et al. 1993, Kwok et al. 1994, Liang et al. 1996, Schilling et al. 1996, Welp et al. 1996\] and BSCCO \[Pastoriza et al. 1993, Zeldov et al. 1995, Hanaguri et al. 1996\]. To estimate the melting temperature one may use the phenomenological Lindemann criterion $`𝐮^2^{1/2}=c_{\mathrm{Li}}a_{\mathrm{}},`$ (2.21) where $`𝐮`$ denotes the displacement of a vortex line from its rest position, $`\mathrm{}`$ the thermal average and $`c_{\mathrm{Li}}0.1\mathrm{}0.2`$ is the Lindemann number. Since the shear modulus $`c_{66}`$ vanishes at $`H_{\mathrm{c1}}^{\mathrm{MF}}`$ and $`H_{\mathrm{c2}}^{\mathrm{MF}}`$, melting will occur by approaching both critical fields. Close to $`H_{\mathrm{c1}}`$ the melting line $`H_{\mathrm{m1}}(T)`$ is roughly given by \[Fisher et al. 1991a\] $`{\displaystyle \frac{H_{\mathrm{m1}}H_{\mathrm{c1}}}{H_{\mathrm{c1}}}}\left({\displaystyle \frac{\lambda }{\mathrm{\Lambda }_T}}\right)^2.`$ (2.22) The region $`H_{\mathrm{c1}}<H<H_{\mathrm{m1}}`$, where the vortex lines form a liquid, is extremely small, except for the vicinity of $`T_\mathrm{c}(H=0)`$ where $`\lambda `$ diverges. $`H_{\mathrm{c1}}`$ is reduced with respect to $`H_{\mathrm{c1}}^{\mathrm{MF}}`$ due to fluctuations \[Nelson 1988, Nelson and Seung 1989\]. We note, however, that for a proper calculation of the melting curves the dispersion of the elastic constants has to be taken into account. In anisotropic and layered superconductors, \[Blatter and Geshkenbein 1996\] (?), following an earlier suggestion by \[Brandt et al. 1996\] (?), found an additional fluctuation-induced attractive van-der-Waals interaction between vortex lines, which may lead at very low temperatures to a first order transition between the Meissner and the Abrikosov lattice phase. At large fields the melting line $`H_{\mathrm{m2}}(T)`$ is reached if \[Brandt 1989, Houghton et al. 1989\] $`a_{\mathrm{}}(H_{\mathrm{m2}}){\displaystyle \frac{\lambda ^2}{4c_{\mathrm{Li}}^2\mathrm{\Lambda }_T}}.`$ (2.23) For $`H_{\mathrm{m1}}<H<H_{\mathrm{m2}}`$ the vortex lines form a solid (cf. figure 1). Alternatively, one could start directly from the GL-Hamiltonian and study the fluctuation corrections in the vicinity of $`H_{\mathrm{c2}}^{\mathrm{MF}}`$ (see, e.g., \[Lee and Shenoy 1972\] (?), \[Bray 1974\] (?), \[Thouless 1975\] (?), \[Ruggeri and Thouless 1976\] (?), \[Ruggeri 1979\] (?), \[Brézin et al. 1985\] (?), \[Affleck and Brézin 1985\] (?), \[Brézin et al. 1990a\] (?), \[Radzihovsky 1995b\] (?)). As first observed by \[Lee and Shenoy 1972\] (?), the fluctuations in a $`d`$-dimensional superconductor in an external field are like those of a $`(d2)`$-dimensional system in zero field, suggesting that the mean-field phase transition to an Abrikosov phase with ODLRO is destroyed by fluctuations in dimensions $`d<d_{\mathrm{cl}}^{\mathrm{ODLRO}}=4`$ (the upper critical dimension of the system is now $`d_{\mathrm{cu}}^{\mathrm{ODLRO}}=6`$). This somewhat surprising conclusion is in agreement with calculations of Moore (?, ?) and others (\[Glazman and Koshelev 1991a\] (?, ?), \[Ikeda et al. 1992\] (?)), who found the destruction of ODLRO by strong phase fluctuations below a lower critical dimension, $`dd_{\mathrm{cl}}^{\mathrm{ODLRO}}`$, starting from the existence of a periodic solution, i.e. of a vortex lattice within the GL-theory. Detailed considerations show that the lower critical dimension $`d_{\mathrm{cl}}^{\mathrm{ODLRO}}=4`$ for systems with screening and $`d_{\mathrm{cl}}^{\mathrm{ODLRO}}=3`$ for systems without screening (Moore ?, ?). Since the aforementioned calculations on fluctuation effects close to $`H_{\mathrm{c2}}^{\mathrm{MF}}`$ use the LLL approximation which usually neglects screening, it is clear that a simple dimensionality shift by 2 does not work here. Qualitatively, it is plausible that screening as an additional source of fluctuations increases $`d_{\mathrm{cl}}^{\mathrm{ODLRO}}`$. A similar effect is also observed for gauge glasses (see section 2.4). Quantitatively, the shift of $`d_{\mathrm{cl}}^{\mathrm{ODLRO}}`$ can be traced back to the strong dispersion of the tilt modulus for $`k>\lambda ^1`$ \[Moore 1992\]. (Note that the compression modulus does not affect the formation of ODLRO.) Breakdown of dimensionality reduction by 2 was also shown by \[Radzihovsky and Balents 1996\] (?) in a layered superconductor in a parallel field, where $`d_{\mathrm{cl}}^{\mathrm{ODLRO}}=2.5`$ but the upper critical dimension $`d_{\mathrm{cu}}^{\mathrm{ODLRO}}=5`$, and the vortex lattice is still stable in $`d2`$ dimensions. ODLRO is conventionally defined by a non-vanishing limit of the gauge invariant pair correlation function $`C_2(𝐫_1,𝐫_2;\mathrm{\Gamma })=\mathrm{\Psi }(𝐫_1)\mathrm{\Psi }^{}(𝐫_2)e^{i(2\pi /\mathrm{\Phi }_0)_\mathrm{\Gamma }𝑑𝐫𝐀}`$ (2.24) for $`|𝐫_1𝐫_2|\mathrm{}`$. ($``$ denotes the average over thermal fluctuations.) Note that $`C_2(𝐫_1,𝐫_2;\mathrm{\Gamma })`$ itself depends on the path $`\mathrm{\Gamma }`$ between $`𝐫_1`$ and $`𝐫_2`$ along which the vector potential is integrated. The proposal of \[Moore 1992\] (?) to keep only the longitudinal component of $`𝐀`$ to make $`C_2`$ path independent but preserve its gauge invariance corresponds in the London gauge to the complete neglection of the phase factor $`e^{i(2\pi /\mathrm{\Phi }_0)_\mathrm{\Gamma }𝑑𝐫𝐀}`$ in equation (2.24). In fact, a non-vanishing asymptotic expression for $`C_2(𝐫_1,𝐫_2;\mathrm{\Gamma })`$ if $`|𝐫_1𝐫_2|\mathrm{}`$ may not be the appropriate definition for the existence of long-range order in cases in which topological defects are forced into the system by external boundary conditions or (as in type-II superconductors) external fields. The most simple counter example is an Ising magnet below $`T_c`$ with anti-periodic boundary conditions, which force a domain wall into the system. Wall fluctuations will then suppress magnetic correlations. The loss of ODLRO due to thermal fluctuations in type-II superconductors – if concluded from the asymptotic behaviour of $`C_2`$ – is related to the fact that phase fluctuations $`\delta \varphi (𝐫)`$ of the order parameter are related to (shear) distortions $`𝐮(𝐫)`$ of the vortex-line lattice via \[Moore 1992, Ikeda et al. 1992, O’Neill and Moore 1993, Baym 1995\] $$\delta \varphi (𝐤)=\frac{2\pi i}{a^2}\frac{k_xu_y(𝐤)k_yu_x(𝐤)}{k^2}.$$ (2.25) Here we used the Fourier transforms $`\delta \varphi (𝐤)`$ and $`𝐮(𝐤)`$ of the fields $`\delta \varphi (𝐫)=\varphi (𝐫)\varphi _0(𝐫)`$ and $`𝐮(𝐫)`$ and $`𝐤=(k_x,k_y,𝐤_{})`$. $`\varphi _0(𝐫)`$ denotes the phase in the ground state. Note that equation (2.25) is not in contradiction to equation (2.10) since $`\mathbf{}^2\varphi _\mathrm{v}(𝐫)0`$ only at the vortex sites. Equation (2.25) implies that on large length scales phase fluctuations are amplified displacement fluctuations of the vortex-line lattice. Destruction of ODLRO below $`d=4`$ dimensions does not mean however – this is the point of view we will take here – the absence of any ordered phase. This can be concluded from the London approximation (2.15): Expressing $`𝐁(𝐫)`$ with the help of (2.14) in terms of the vortex degrees of freedom, the only phase fluctuations left are spin-wave like which lead to a destruction of conventional long-range order only in $`d2`$ dimensions. The energy of the vortex lattice can be approximated at low $`T`$ by an elastic Hamiltonian (2.16) which may be supplemented by a contribution from dislocations. For $`d3`$ such a system shows a phase with translational long-range order (TLRO) for $`H_{\mathrm{m1}}(T)<H<H_{\mathrm{m2}}(T)`$. (In $`d=2`$ dimensions, elastic fluctuations reduce the order to quasi-TLRO characterized by a power-law decay of correlations.) In the elastic description, which we will use in most parts of this review, the appropriate order parameter for TLRO is $$\psi _𝐐(𝐫)=e^{i𝐐𝐮(𝐫)}$$ (2.26) with $`\psi _𝐐(𝐫)0`$ in the Abrikosov phase. $`𝐐`$ denotes a reciprocal lattice vector of the Abrikosov lattice. From this point of view the loss of ODLRO has no physical significance. $`C_2`$ is therefore not the appropriate quantity to define order. This conclusion is in agreement with a number of numerical investigations in which a clear indication for a transition to an ordered phase was seen in two and three dimensions. \[Hu and MacDonald 1993\] (?) and \[Kato and Nagaosa 1993a\] (?) have considered the pair correlation function of the superfluid density $`C_4(𝐫_1,𝐫_2)=|\mathrm{\Psi }(𝐫_1)|^2|\mathrm{\Psi }(𝐫_2)|^2.`$ (2.27) Using the LLL approximation both groups found a clear indication for a first order melting transition in $`d=2`$. Their work was extended by \[Šašik and Stroud 1993\] (?, ?) to three-dimensional systems. They considered the helicity modulus $`\gamma _{\alpha \beta }(𝐫,𝐫^{})={\displaystyle \frac{j_\alpha (𝐫)}{A_\beta ^{}(𝐫^{})}}|_{𝐀^{}=\mathrm{𝟎}}={\displaystyle \frac{c}{4\pi \lambda ^2}}\delta (𝐫𝐫^{})\delta _{\alpha \beta }{\displaystyle \frac{1}{cT}}j_\alpha (𝐫)j_\beta (𝐫^{})_c,`$ (2.28) where $`𝐀^{}`$ is an additional vector potential and $`𝐣`$ the current density. In the LLL (where $`\lambda \mathrm{}`$) they found a rapid increase of $`\gamma _{zz}`$ at the freezing temperature, indicating the formation of a vortex-line lattice. \[Šašik et al. 1995\] (?) have subsequently shown that the vortex liquid-to-solid transition is not accompanied by a divergence of the correlation length for phase coherence. In their simulation $`C_2`$ decays exponentially even in the solid phase in agreement with the predictions of \[Moore 1992\] (?), \[Ikeda et al. 1992\] (?), and \[Baym 1995\] (?). It has to be mentioned, however, that the point of view we adopt in this article – namely that the loss of ODLRO does not rule out the existence of a vortex lattice – is not shared by all authors. In particular Moore (?, ?, ?) argues that there is no mixed phase in type-II superconductors at finite temperatures. The observed effects in the behaviour of resistance and magnetization are explained as cross-over phenomena which would disappear in the thermodynamic limit. Moreover, the vortex lattices found in the Monte-Carlo simulations mentioned before is considered to be an artifact produced by the quasi-periodic boundary conditions used in these simulations. Moore and co-workers \[O’Neill and Moore 1992, O’Neill and Moore 1993, Lee and Moore 1994, Dodgson and Moore 1997, Kienappel and Moore 1999, Moore and Pérez-Garrido 1999\] have tried in their own Monte-Carlo simulations to avoid these effects (which they consider to be crucial) by placing the two-dimensional superconductor on the surface of a sphere. In their studies no freezing transition to a vortex-lattice state was observed. It should however be mentioned that zero-energy modes connected with the rigid rotations of the film on the sphere and disclinations arising from topological constraints for a triangular lattice on the sphere may obscure the transition. To conclude: Although the absence of ODLRO at finite temperatures in the mixed phase of type-II superconductors (see, e.g., \[Houghton et al. 1990\]) and its consequences for the existence of a vortex lattice are still under debate, the most simple scenario is the existence of a vortex lattice in the absence of ODLRO below $`d=4`$ dimensions. The continuous translational invariance of the system is reduced to finite translations by lattice vectors $`𝐗`$ of the Abrikosov lattice. In contrast to ODLRO, the lower critical dimension for TLRO in pure systems is $`d_{\mathrm{cl}}^{\mathrm{TLRO}}=2`$. Since the critical regime is enlarged in a non-zero external field (in $`d=3`$) to \[Ikeda et al. 1989\] $`{\displaystyle \frac{T_\mathrm{c}(H)T}{T_\mathrm{c}(H)}}\mathrm{Gi}^{1/3}\left({\displaystyle \frac{H_{\mathrm{c2}}^{\mathrm{MF}}(T)}{H_{\mathrm{c2}}^{\mathrm{MF}}(0)}}\right)^{2/3}`$ (2.29) with respect to the zero-field critical region, but is still too small to describe the difference between $`H_{\mathrm{c2}}^{\mathrm{MF}}(T)`$ and the melting line, \[Feigelman et al. 1993\] (?) have argued that between the Abrikosov lattice and the normal phase there may be an intermediate liquid phase which still shows longitudinal superconductivity. We will not follow this idea here since more recent extensive simulations show only a single transition between the Abrikosov and the normal phase \[Hu et al. 1997, Hu et al. 1998, Nguyen and Sudbø 1998, Nordborg and Blatter 1997, Nordborg and Blatter 1998, Olsson and Teitel 1999\], although further proposals for an intermediate phase were made recently \[Tešanović 1999, Nguyen and Sudbø 1999\]. In the rest of this article we will ignore the existence of these exotic phases in pure superconductors (although their existence cannot be ruled out completely), but concentrate instead on the influence of randomly distributed frozen impurities on the vortex-line lattice phase. The impurities will be assumed to be completely uncorrelated as already mentioned in the introduction. Columnar or planar defects lead to very different physics (for a recent review on the so-called Bose glass, see \[Täuber and Nelson 1997\] (?) and references therein) and will not be discussed in this paper. ### 2.4 The influence of disorder Disorder can be introduced in the Ginzburg-Landau model (2.1) by assuming small random contributions to all parameters which characterize the system, i.e. $`\alpha ,\beta ,m`$ and $`𝐇`$. We will restrict ourselves here to the case of systems with random mean-field transition temperature $`T_{\mathrm{c0}}`$, i.e. we substitute $$T_{\mathrm{c0}}T_{\mathrm{c0}}+\delta T_{\mathrm{c0}}(𝐫)$$ (2.30) with $$\overline{\delta T_{\mathrm{c0}}(𝐫)\delta T_{\mathrm{c0}}(𝐫^{})}=\xi ^2\delta _\xi (𝐫𝐫^{})\delta T_{\mathrm{c0}}^2$$ (2.31) where the overbar denotes the disorder average and $`\delta _\xi (𝐫)`$ is a $`\delta `$-function of widths $`\xi `$. In mean-field theory, a randomness in $`T_{\mathrm{c0}}`$ would correspond to a smearing out of the transition. However, as first shown by \[Khmelnitskii 1975\] (?), thermal fluctuations permit the occurrence of a sharp phase transition. According to the so-called Harris criterion \[Harris 1974\] weak randomness does not change the critical behaviour if the exponent of the specific heat of the pure system $`\alpha _{\mathrm{pure}}`$ is negative, or, equivalently, if the correlation length exponent $`\nu _{\mathrm{pure}}`$ obeys $`\nu >2/d`$. Since for the $`XY`$-model in three dimensions $`\alpha `$ is negative \[Lipa et al. 1996\], the zero-field critical behaviour for type-II superconductors should be unchanged in the outer critical region. This would also apply in case that behaviour of the pure system in the inner critical region is also of (inverted) $`XY`$-type. On the contrary, for type-I superconductors it was shown by \[Boyanovsky and Cardy 1982\] (?) that a sufficiently large amount of disorder may convert the first order transition into a second order one. If an external field is applied, the situation is quite different. Because of the problems connected with ODLRO even in pure systems, which were discussed in the previous chapter, it seems to be expedient to first consider the influence of disorder on the structural properties of the mixed phase. Inside the Abrikosov phase, disorder leads to a destruction of translational long-range order as first shown by \[Larkin 1970\] (?). This follows from the fact that a randomness in the local value of $`T_c(𝐫)`$ leads to a random potential acting on the vortices. The specific form of the resulting coupling of the disorder to the vortex displacements will be given in the following chapters. The disorder averaged order parameter $`\overline{e^{i𝐐𝐮}}`$ for TLRO vanishes since disorder leads to diverging displacement $`𝐮`$ in the limit of large system sizes. However, it will be shown below that the correlation function $$S(𝐐,𝐫)=\overline{e^{i𝐐[𝐮(𝐫)𝐮(\mathrm{𝟎})]}}$$ (2.32) may still obey an algebraic decay, which shows up in Bragg peaks in the structure factor (Giamarchi and Le Doussal ?, ?, \[Emig et al. 1999\]). Moreover, in analogy with spin glass theory (see \[Binder and Young 1986\] (?)) one may consider the positional glass correlation function $$S_{\mathrm{PG}}(𝐐,𝐫)=\overline{\left|e^{i𝐐[𝐮(𝐫)𝐮(\mathrm{𝟎})]}\right|^2}$$ (2.33) as another signature of the existence of some residual, ‘glassy’ order of the Abrikosov lattice. Below we will call a system a positional vortex glass if $`S_{\mathrm{PG}}(𝐐,𝐫)`$ decays not faster than a power law for $`|𝐫|\mathrm{}`$. For $`T0`$, $`S_{\mathrm{PG}}(𝐐,𝐫)`$ approaches one. At non-zero temperature, (2.33) measures the strength of thermal fluctuations of the vortex lines around the ground state. Two limiting cases seem to be conceivable: If the disorder acts effectively as a random force on the vortex lines, then the thermal fluctuations around the disordered ground state are identical to those of the pure system. In this case (2.33) is non-zero above $`d=2`$ dimensions. In the opposite case of very strong pinning the vortex lines can be considered to fluctuate thermally in a kind of narrow parabolic potential and (2.33) may be finite even in $`d<2`$. The relevance of these structural correlation functions for the glassy behaviour of the mixed phase will be further discussed in the following chapters. A complementary discussion of the glassy behaviour was proposed by M. P. A. Fisher (?) and \[Fisher et al. 1991a\] (?). These authors use the correlation function for ODLRO as the starting point. Clearly, because of the relation (2.25), $`\overline{\mathrm{\Psi }(𝐫)\mathrm{\Psi }^{}(\mathrm{𝟎})}`$ will vanish for $`r\mathrm{}`$ in $`d<4`$ dimensions. In analogy to spin glasses they therefore introduce the correlation function for phase-coherent vortex-glass order (using the London gauge $`\mathbf{}𝐀=0`$): $$C_{\mathrm{VG}}(𝐫)=\overline{\left|<\mathrm{\Psi }^{}(𝐫)\mathrm{\Psi }(\mathrm{𝟎})>\right|^2}.$$ (2.34) It is instructive to consider (2.34) in the London limit where $`C_{\mathrm{VG}}(𝐫)|\mathrm{\Psi }_0|^2\overline{|e^{i[\delta \varphi (𝐫)\delta \varphi (\mathrm{𝟎})]}|^2}`$. $`\delta \varphi (𝐫)`$ describes the fluctuations around the ground state pattern $`\varphi _0(𝐫)`$. If the disorder is of random-force type, according to (2.25) thermal fluctuations should destroy $`C_{\mathrm{VG}}(𝐫)`$ below $`d=4`$. On the other hand for thermal fluctuations in a parabolic potential around the ground state, $`C_{\mathrm{VG}}(𝐫)`$ should be finite. \[Dorsey et al. 1992\] (?) and Ikeda (?, ?) have calculated the vortex glass susceptibility $`\chi _{\mathrm{VG}}=d^drC_{\mathrm{VG}}(𝐫)`$ for a GL-Model with random transition temperature in the LLL approximation. \[Dorsey et al. 1992\] (?) found in a mean-field calculation a second-order transition with a diverging vortex-glass susceptibility by approaching the vortex-glass transition temperature $`T_\mathrm{g}(H)`$, which is slightly below $`T_{\mathrm{c2}}(H)`$. In mean-field theory, the vortex-glass correlation length $`\xi _{\mathrm{VG}}`$ diverges with a mean-field exponent $`\nu _{\mathrm{VG}}^{\mathrm{MF}}=1/2`$. From this one can expect the existence of a non-zero limit of $`C_{\mathrm{VG}}(𝐫)`$ for $`|𝐫|\mathrm{}`$ below $`T_\mathrm{g}(H)`$. Taking critical fluctuations into account, which can be done within a $`d=6ϵ`$ expansion, the model is found to be in the some universality class as the Ising spin glass. In particular, the dynamical critical exponent is found to be $`z=2(2\eta )`$, where $`\eta ϵ/6`$. It should be observed that a similarity between the Ising spin glass and vortex glasses was also mentioned for so-called gauge-glass models (see \[Huse and Seung 1990\] (?)). On the other hand, the results of \[Dorsey et al. 1992\] (?) and Ikeda (?, ?) – if applied to $`d=3`$ dimensions – cannot be so easily reconciled with the findings from the elastic description of a vortex lattice in a random potential. As we will discuss in chapter 6, the vortex lattice exhibits indeed a glassy phase (the Bragg glass) for $`2<d4`$, in which the correlation function $`C_{\mathrm{VG}}(𝐫)`$, if calculated to lowest order in $`ϵ=4d`$, vanishes for large $`|𝐫|`$ exponentially. The reason for this decay consists in the strong thermal fluctuations of the phases around the (distorted) ground state. To order $`ϵ`$, these phase fluctuations are only weakly suppressed by disorder and hence $`C_{\mathrm{VG}}(𝐫)`$ decays exponentially. However, higher order terms in $`ϵ`$ may still change this result. In principle, there could be two glassy phases, which show a non-vanishing $`S_{\mathrm{PG}}(𝐐,𝐫)`$ and $`C_{\mathrm{VG}}`$, respectively. For the moment we consider it to be more likely that there is only one glassy phase and the discrepancy between the results follows from the use of different approximations valid in $`d=6ϵ`$ and $`d=4ϵ`$ dimensions, respectively. So far we assumed that the disorder is weak. On the other hand gauge-glass like models were proposed for the description of granular superconductors or systems with strong disorder \[Ebner and Stroud 1985, John and Lubensky 1986\]. Each superconducting grain of centre position $`𝐫_i`$ is described by the phase $`\varphi _i`$ of the order parameter which is assumed to be constant within a grain. The Hamiltonian then reads (see e.g. \[Wengel and Young 1996\] (?)) $$=J\underset{<ij>}{}\mathrm{cos}(\varphi _i\varphi _jA_{ij}\lambda _0^1a_{ij})+\frac{1}{2}\underset{\mathrm{}}{}(\mathbf{}𝐚)^2,$$ (2.35) where $`a_{ij}=\underset{𝐫_j}{\overset{𝐫_i}{}}𝐚(𝐫)𝑑𝐫`$. $`\underset{<ij>}{}`$ is the sum of all nearest neighbors of a cubic lattice and $`\underset{\mathrm{}}{}`$ is the sum over all plaquettes. $`𝐚`$ denotes the fluctuations of the vector potential which are limited by the bare screening length $`\lambda _0`$. The influence of the external field as well as the contribution from randomness are assumed to be included in the $`A_{ij}`$ which are taken to be independent random variables with a distribution between 0 and $`2\pi `$. A detailed discussion of the relation between the vortex glass (the expression is here understood in the sense that it describes the glassy phase of an impure type-II superconductor in an external field) and gauge glasses is given in \[Blatter et al. 1994\] (?). The model (2.35) is in particular isotropic in contrast to the GL-Hamiltonian which shows a pronounced anisotropy due to the presence of the external field $`𝐇`$. In addition, the disorder of the gauge glass has a nature which is completely different from the one in equation (2.30). The first one couples to the vortices via the phase of the order parameter, while the latter one couples only via the amplitude. As a consequence, the gauge-glass disorder distorts the vortices much more than local variations of $`T_{\mathrm{c0}}`$. For $`\lambda _0\mathrm{}`$, i.e. in the absence of screening, the gauge glass was investigated numerically by a number of authors (see \[Fisher et al. 1991b\] (?), \[Gingras 1992\] (?), \[Cieplak et al. 1991\] (?), \[Reger et al. 1991\] (?), \[Cieplak et al. 1992\] (?), \[Hyman et al. 1995\] (?), \[Maucourt and Grempel 1998\] (?), \[Kosterlitz and Akino 1998\] (?), \[Olson and Young 1999\] (?)). While in two dimensions a transition to a glass phase is found to be only at $`T=0`$, this transition takes place at finite temperatures in three dimensions. \[Huse and Seung 1990\] (?) found at the transition a diverging gauge-glass susceptibility $$\chi _{\mathrm{GG}}=\underset{j}{}C_{\mathrm{GG}}(𝐫_i𝐫_j),$$ (2.36) where $`C_{\mathrm{GG}}`$, in analogy with (2.34), denotes the gauge glass correlation function $$C_{\mathrm{GG}}(𝐫_i𝐫_j)=\overline{\left|e^{i(\varphi _i\varphi _j)}\right|^2}.$$ (2.37) The divergence of $`\chi _{\mathrm{GG}}`$ may signal the transition to a phase with non-zero limit of $`C_{\mathrm{GG}}(𝐫)`$ for $`|𝐫|\mathrm{}`$. If one assumes (2.34) as the definition of vortex glass order, then the gauge glass would be a vortex glass. More recently, the case of a finite $`\lambda _0`$ was considered \[Bokil and Young 1995, Wengel and Young 1996, Wengel and Young 1997, Kisker and Rieger 1998\]. It turns out that screening seems to destroy the gauge glass transition in three dimensions. More recently \[Kawamura 1999\] (?) and \[Pfeiffer and Rieger 1999\] (?) have attempted to include the effect of anisotropy into the gauge glass model by assuming an extra contribution to $`A_{ij}`$ arising from the external field. However, also in this case no finite temperature glass transition was found. ## 3 Directed elastic manifolds in a random potential In the course of this article we will restrict our consideration of fluctuations to vortices and we will ignore other independent fluctuations such as those of the condensate amplitude or of the magnetic field. Vortices will be treated within the London picture, where vortex lines are represented as string-like objects. For low temperatures and weak disorder the vortex lines will fluctuate only weakly around the ground state of the vortex-line lattice (VLL), the Abrikosov lattice, where all lines are directed along the magnetic field. Before we analyze the VLL as an ensemble of vortex lines in its full complexity it is instructive to study a single vortex line. In general, a vortex can be characterized by two dimensions which depend on the physical realization under consideration: the ‘internal’ dimension $`D`$ of the vortex considered as a ‘directed manifold’, and the number $`N`$ of its displacement components. For example, a vortex line in a bulk superconductor has $`(D,N)=(1,2)`$, a single vortex line in a superconducting film has $`(D,N)=(1,1)`$, and a point vortex in a film corresponds to $`(D,N)=(0,2)`$. This concept of elastic manifolds also applies to other physical systems such as interfaces between magnetic domains, for which $`(D,N)=(2,1)`$. In all these examples the spatial dimension of the system is $`d=D+N`$, since displacements are possible only in directions orthogonal to the $`𝐳`$ direction(s) along which the manifold is spanned. It is worthwhile to mention at this point that the analysis of random manifolds applies not only to single vortex lines, but to a certain extent (i.e., within a regime of length scales) also to vortex line lattices. For example, the VLL in a weakly disordered bulk superconductor resembles over a large range of length scales an elastic manifold with $`(D,N)=(3,2)`$. In this case, where the elastic manifold is spanned in all spatial dimensions including the displacement directions, $`d=D`$. Nevertheless, there is a fundamental difference between manifolds and VLLs, which is crucial for the physics on large scales: VLLs have a periodic structure in contrast to manifolds. To parameterize vortex conformations, we use the $`D`$-dimensional coordinate $`𝐳`$ along the direction of the magnetic field and a $`N`$-dimensional coordinate $`𝐱𝐱(𝐳)`$ in transverse directions, see figure 2. In this chapter we are mainly interested in qualitative aspects of vortex fluctuations in the presence of disorder. To this end we describe the elastic energy of a vortex line, which arises from the kinetic energy of the supercurrents and the magnetic field energy, in a harmonic approximation (i.e., we keep only terms of second order in the displacement) and we ignore non-localities or possible anisotropies in the elastic stiffness constant $`\epsilon _{}`$. Then the elastic energy can be written as $`_{\mathrm{el}}`$ $`=`$ $`{\displaystyle d^Dz\frac{\epsilon _{}}{2}(\mathbf{}_{}𝐱)^2}.`$ (3.1) Therein the gradient $`\mathbf{}_{}(\frac{}{z_1},\mathrm{},\frac{}{z_D})`$ acts along the longitudinal $`𝐳`$ directions. Pinning of vortex lines due to the presence of impurities, grain boundaries etc. can be described by a potential $`V(𝐱,𝐳)`$ which is ‘quenched’, i.e., frozen on the relevant time scale for vortex fluctuations. It yields a contribution $`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle d^DzV(𝐱(𝐳),𝐳)}`$ (3.2) to the total energy $`=_{\mathrm{el}}+_{\mathrm{pin}}`$ of a vortex line. For simplicity, we will discuss here only point-like disorder. We will assume that this potential is Gaussian distributed with zero average and variance $`\overline{V(𝐱,𝐳)V(𝐱^{},𝐳^{})}`$ $`=`$ $`\mathrm{\Delta }(𝐱𝐱^{})\delta (𝐳𝐳^{}).`$ (3.3) For simplicity we will restrict ourselves to disorder with short-ranged and isotropic correlations. Then the correlator $`\mathrm{\Delta }(𝐱)=\mathrm{\Delta }(|𝐱|)`$ can be characterized by its integral $`\mathrm{\Delta }_0`$ $``$ $`\widehat{\mathrm{\Delta }}(\mathrm{𝟎})={\displaystyle d^Nx\mathrm{\Delta }(𝐱)}`$ (3.4) and a correlation length. Such a pinning potential can arise from point-like impurities that locally suppress the density of the superconducting condensate. The correlation length of the disorder then also coincides with the superconducting coherence length $`\xi `$. In principle, one can retain a finite correlation length in the $`𝐱`$ and $`𝐳`$ directions. However, it turns out a posteriori that – as long as the correlation length does not exceed the smallest scale of vortex conformations, which also is $`\xi `$ – correlations in $`𝐳`$ direction are irrelevant. Subsequently, we will often write $`\mathrm{\Delta }(𝐱)=\mathrm{\Delta }_0\delta _\xi (𝐱)`$, where $`\delta _\xi `$ denotes a Dirac delta-function smeared out on a scale $`\xi `$. For semi-quantitative purposes we occasionally use $`\delta _\xi (𝐱)=(2\pi \xi ^2)^{N/2}\mathrm{exp}(𝐱^2/2\xi ^2)`$. If we denote the pinning force of an individual impurity by $`f_{\mathrm{pin}}`$, then $`\mathrm{\Delta }_0f_{\mathrm{pin}}^2n_\mathrm{i}^{(d)}\xi ^{2N+2},`$ (3.5) where $`n_\mathrm{i}^{(d)}`$ denotes the concentration of the impurities in the $`d`$ dimensional space. Directed manifolds in disorder represent a paradigm for systems which are dominated by disorder. Since disorder leads to a frustrating competition between elastic and pinning energies, the structural order can be reduced substantially. In addition, the dynamics of the system can get extremely slow due to the quenched nature of disorder. Therefore such a system can be called a ‘glass’. Because of the simplicity of the manifold model, it plays a paradigmatic role for glassy systems and it allows the identification and the understanding of many characteristic features of a more complicated ‘vortex glass’. ### 3.1 Equilibrium properties As we have seen, vortices in impure superconductors are a realization of ‘elastic manifolds’ in ‘random media’ which are paradigms for disordered systems. We now proceed to summarize briefly some key properties of the latter model system. For more detailed presentations about these random manifolds the reader is referred to review articles \[Kardar 1994, Halpin-Healy and Zhang 1995, Lässig 1998a\], in particular to \[Blatter et al. 1994\] (?) for a discussion in the context of vortex systems, as well as to \[Nattermann and Rujan 1989\] (?), \[Belanger and Young 1991\] (?) and \[Young 1998\] (?) in the context of disordered magnets. Subsequently we will focus our attention on the question, in what physical quantities ‘glassiness’ appears in these systems. From a principal point of view, it is important to note that disorder breaks symmetries of the Hamiltonian of the pure system: translation invariance $`𝐱(𝐳)𝐱(𝐳)+𝐱_0`$ for a constant shift $`𝐱_0`$ of vortex lines, as well as an analogous rotation symmetry $`𝐱(𝐳)𝓡𝐱(𝐳)`$, for a rotation matrix $`𝓡`$. Therefore it is interesting to examine to what extent the physical state of the manifold in disorder actually reflects these broken symmetries, or whether these symmetries are restored due to thermal fluctuations. #### 3.1.1 Structure One quantity of primary interest is the structure of the manifold in disorder which can be described in terms of displacement correlations. In the absence of disorder the vortex line has a flat ground state, $`𝐱(𝐳)𝐗`$ for all $`𝐳`$. Introducing the displacement $`𝐮(𝐳)𝐱(𝐳)𝐗`$, shape fluctuations can be described by the relative displacement at points separated by a distance $`𝐳𝐳^{}`$ parallel to the magnetic field: $`W(𝐳𝐳^{})`$ $``$ $`\overline{[𝐮(𝐳)𝐮(𝐳^{})]^2}|𝐳𝐳^{}|^{2\zeta }.`$ (3.6) As already introduced above, $`\mathrm{}`$ and $`\overline{\mathrm{}\text{}}`$ denote the thermal and disorder average respectively. For large distances this quantity typically follows a power law with the roughness exponent $`\zeta =\zeta (D,N)`$. The manifold is called flat, if $`W`$ is finite for $`|𝐳𝐳^{}|\mathrm{}`$ (then the convergence of $`W`$ to its asymptotic value can be described by an exponent $`\zeta <0`$), whereas it is called rough if $`W`$ diverges (i.e., $`\zeta 0`$, where $`\zeta =0`$ typically corresponds to $`W(𝐳)\mathrm{ln}^\alpha z`$ with some power $`\alpha `$). In the absence of disorder thermal fluctuations are described by an exponent $$\zeta _{\mathrm{th}}=\frac{2D}{2}.$$ (3.7) The presence of disorder may increase the roughness and lead to a larger exponent $`\zeta >\zeta _{\mathrm{th}}`$. In particular, disorder always induces roughness in dimensions $`2<D4`$, as we will discuss in more detail below. For $`D<2N/(2+N)`$ disorder is irrelevant at sufficiently high temperatures. There is a phase transition (see e.g. \[Imbrie and Spencer 1988\] (?) for the case $`D=1`$) between a low-temperature phase, where the manifold is disorder dominated and essentially shows the structure of the ground state, and a high-temperature phase, where the manifold is entropically driven out of the ground state and shows a structure as in the absence of disorder. Actually, the physical situation can be more complicated if an upper critical dimension $`N_{\mathrm{cu}}`$ exists, above which the low-temperature phase is governed by the Gaussian exponents. The existence of an upper critical dimension is still controversial. \[Moore et al. 1995\] (?) argue for $`N_{\mathrm{cu}}=4`$ and \[Lässig and Kinzelbach 1997\] (?, ?) argue for $`N_{\mathrm{cu}}4`$ in $`D=1`$. We will not further discuss this possible complication here, assuming $`N_{\mathrm{cu}}`$ to be large enough such that the vortex systems of physical interest are not concerned. For $`D2N/(2+N)`$ one can think of the manifold as having a unique disorder-dominated ground state. For all temperatures entropic effects are too weak to detach the manifold from its ground state. 3.1.1a Structural order parameter In order to have a tool to quantify the structural order of the manifold, we introduce $`\psi _𝐤(𝐳)`$ $``$ $`e^{i𝐤𝐮(𝐳)}`$ (3.8) as order parameter. If the manifold performs only weak thermal fluctuations around its ground state, $`\psi _𝐤(𝐳)0`$ and disorder actually breaks the translation symmetry of the manifold. Otherwise, if thermal fluctuations detach the manifold from its ground state, $`\psi _𝐤(𝐳)=0`$. The correlation function of this order parameter, $`S(𝐤,𝐳𝐳^{})\overline{\psi _𝐤^{}(𝐳)\psi _𝐤(𝐳^{})}\mathrm{exp}\left({\displaystyle \frac{1}{2}}𝐤𝐖(𝐳𝐳^{})𝐤\right),`$ (3.9) is related to the displacement correlation $`W_{\alpha \beta }(𝐳𝐳^{})`$ $``$ $`\overline{[u_\alpha (𝐳)u_\alpha (𝐳^{})][u_\beta (𝐳)u_\beta (𝐳^{})]}.`$ (3.10) The function $`W`$, previously introduced in (3.6), is simply the trace of the matrix $`W_{\alpha \beta }`$. The approximate relation in (3.9) neglects higher cumulants of the displacement distribution and holds in general only for small displacement fluctuations. It even holds for large displacement fluctuations and actually is an identity, if the fluctuations of $`𝐮`$ have Gaussian distribution. 3.1.1b Perturbative analysis A first qualitative insight into the relevance of disorder can be obtained from an elementary perturbative analysis. Such an analysis was performed originally by \[Larkin 1970\] (?) for vortex lattices and by \[Efetov and Larkin 1977\] (?) for the closely related charge-density waves. In this approach the manifold (at $`T=0`$) is considered in an absolutely flat reference state $`𝐱(𝐳)=𝐗`$, for which the pinning force $`𝐅^{\mathrm{pin}}(𝐳)\mathbf{}_{}V(𝐱(𝐳),𝐳)`$ is calculated. We denote by $`\mathbf{}_{}(\frac{}{x_1},\mathrm{},\frac{}{x_N})`$ gradients in $`𝐱`$ directions in contrast to $`\mathbf{}_{}`$ for $`𝐳`$ directions. From this force the disorder induced manifold displacement is obtained in linear response theory as $`u_\alpha (𝐪)d^Dze^{i𝐪𝐳}u_\alpha (𝐳)=G_{\alpha \beta }(𝐪)F_\beta ^{\mathrm{pin}}(𝐪)`$ using the response function $`G_{\alpha \beta }(𝐳𝐳^{}){\displaystyle \frac{\delta u_\alpha (𝐳)}{\delta F_\beta ^{\mathrm{pin}}(𝐳^{})}},G_{\alpha \beta }(𝐪)={\displaystyle \frac{1}{\epsilon _{}q^2}}\delta _{\alpha \beta }`$ (3.11) of the free manifold. Thus the displacement correlations are given in linear response by $`\overline{u_\alpha (𝐪)u_\beta (𝐪)}=G_{\alpha \gamma }(𝐪)\overline{F_\gamma ^{\mathrm{pin}}(𝐪)F_\delta ^{\mathrm{pin}}(𝐪)}G_{\beta \delta }(𝐪)={\displaystyle \frac{\mathrm{\Delta }^{(2)}}{\epsilon _{}^2q^4}}\delta _{\alpha \beta },`$ (3.12) where we introduced the variance of the pinning force $`\mathrm{\Delta }^{(2)}`$ via $`\mathrm{\Delta }^{(2)}\delta _{\alpha \beta }_\alpha _\beta \mathrm{\Delta }(𝐱)|_{𝐱=\mathrm{𝟎}}.`$ (3.13) Its value is related to the variance and correlation length of the pinning potential approximately through $`\mathrm{\Delta }^{(2)}={\displaystyle \frac{\mathrm{\Delta }_0}{(2\pi )^{N/2}\xi ^{2+N}}}.`$ (3.14) Here we have assumed a Gaussian form for $`\mathrm{\Delta }(𝐱)`$. The perturbative correlation function (3.12) is characterized by a roughness exponent $$\zeta _{\mathrm{rf}}=\frac{4D}{2},$$ (3.15) which we refer to as the ‘random force’ value. This exponent characterizes the actual correlation function only on sufficiently small scales $`|𝐳|L_\xi `$ below the Larkin length $`L_\xi `$ \[Larkin 1970\], since the perturbative treatment is justified only as long as $`W(𝐳)\xi ^2`$. The Larkin length can be estimated by equating the elastic energy $`E_{\mathrm{el}}\epsilon _{}L_\xi ^{D2}\xi ^2`$ with the pinning energy in the random-force approximation $`E_{\mathrm{pin}}(L^D\xi ^2\mathrm{\Delta }^{(2)})^{1/2}`$ as \[Larkin 1970, Bruinsma and Aeppli 1984\] $`L_\xi [\epsilon _{}^2\xi ^2/\mathrm{\Delta }^{(2)}]^{1/ϵ},`$ (3.16) where $`ϵ4D`$ (to be distinguished from $`\epsilon _{}`$). From this rough analysis the manifold is found to be flat in $`D>4`$ (since $`\zeta _{\mathrm{rf}}<0`$), logarithmically rough in $`D=4`$ (with $`\zeta _{\mathrm{rf}}=0`$), and rough with an exponent $`\zeta _{\mathrm{rf}}>0`$ in $`D<4`$. Although the perturbative analysis can provide a good approximation only on small length scales $`|𝐳|L_\xi `$ (however, in certain cases intermittency can be relevant on small scales \[Bouchaud et al. 1995\]), the roughness of the manifold in $`D4`$ persists in more sophisticated approaches (such as a self-consistent or renormalization-group analysis), which go beyond a perturbative approach. Thus, a single vortex line ($`D=1`$) in a bulk superconductor ($`d=3`$), which is roughened by pure thermal fluctuations (cf. equation (3.7)), is also roughened by disorder at $`T=0`$. In contrast, a VLL in a bulk superconductor, which can be considered as a manifold with $`D=d`$, is not roughened by pure thermal fluctuations but by pinning. 3.1.1c Flory analysis As already mentioned, the above perturbative analysis breaks down on length scales $`|𝐳|L_\xi `$ because perturbation theory does not adequately treat a system with many minima in the potential energy (for a pedagogical example see \[Villain and Séméria 1983\] (?)). Although perturbation theory could be continued to higher orders, it will never be able to describe the displacements on largest scales $`|𝐳|\mathrm{}`$, where the multi-stability (existence of many local minima) of the potential energy landscape is crucial. For the further analysis it is convenient to start from the replica Hamiltonian \[Edwards and Anderson 1975\] $`_n`$ $`=`$ $`_{\mathrm{el},n}+_{\mathrm{pin},n}`$ (3.17) $`=`$ $`{\displaystyle \underset{a,b=1}{\overset{n}{}}}{\displaystyle d^Dz\left\{\frac{\epsilon _{}}{2}\delta ^{ab}(\mathbf{}_𝐳𝐮^a)^2\frac{1}{2T}\mathrm{\Delta }(𝐮^a(𝐳)𝐮^b(𝐳))\right\}},`$ which is obtained from replicating the original system $`n`$ times and performing a disorder average. Note that Greek lower indices denote transverse spatial components, whereas Roman upper indices denote replicas. The analysis of this Hamiltonian is highly non-trivial, since the displacement enters the argument of the disorder correlator $`\mathrm{\Delta }`$. A dimensional analysis shows that $`\mathrm{\Delta }`$ has to be retained in its full functional form and may not be represented by a truncated Taylor expansion in $`D4`$ \[Balents and Fisher 1993\]. Physically, the structure on large scales is determined by a competition between elastic and pinning energies. The Flory argument \[Imry and Ma 1975, Kardar 1987, Nattermann 1987\] allows one to obtain an improved value for the roughness exponent $`\zeta `$ by requiring both energy contributions to scale with the same exponent on large scales $`|𝐳|`$. To be more precise, we rescale $`𝐳=L𝐳^{}`$ and $`𝐮(𝐳)=L^\zeta 𝐮^{}(𝐳^{})`$, where $`L`$ denotes the (variable) length scale on which we consider the system. Then $`_{\mathrm{el},n}L^\theta `$ with an energy scaling exponent $`\theta =2\zeta +D2.`$ (3.18) The fluctuations of $`𝐮^{}`$ are determined by an effective temperature $`T^{}`$, which has to be rescaled like the Hamiltonian, $`T=T^{}L^\theta ,`$ (3.19) in order to keep the Boltzmann factor $`e^{/T}`$ scale invariant. In pure systems it is possible to achieve scale invariance not only of the Boltzmann factor $`e^{_{\mathrm{el}}/T}`$ but of both $`cH_{\mathrm{el}}`$ and $`T`$ by the choice $`\zeta =\zeta _{\mathrm{th}}\frac{2D}{2}`$ and $`\theta _{\mathrm{th}}=0`$. At zero temperature, weak disorder is a relevant perturbation in $`D4`$. Since a short-ranged disorder correlator should essentially scale like an $`N`$-dimensional $`\delta `$-function, the replicated pinning energy scales as $`_{\mathrm{pin}}/TL^{DN\zeta 2\theta }/T_{}^{}{}_{}{}^{2}`$. The exponent of the Boltzmann factor includes now two terms, which after rescaling behave as $`_{\mathrm{el}}/T_{\mathrm{el}}^{}/T^{}=O(1)`$ and $`_{\mathrm{pin}}/T=O(L^{DN\zeta 2\theta })`$. In the limit of vanishing temperatures, a finite width of the distribution of $`\{𝐮^{}\}`$ is only possible if $`L^{DN\zeta 2\theta }=O(1)`$, i.e., if $$\zeta =\zeta _\mathrm{F}\frac{4D}{4+N},$$ (3.20) which implies $`\theta =2\zeta +D2=\frac{2+N}{4+N}(DD_N)`$, where $`D_N\frac{2N}{2+N}`$. Although $`\zeta _\mathrm{F}`$ is an improvement over $`\zeta _{\mathrm{rf}}`$, this result is (in general) not exact, since the scaling behaviour of $`_{\mathrm{pin}}`$ was over-simplified. On the other hand, at finite temperatures $`e^{_{\mathrm{pin},n}/T}`$ is a relevant perturbation to the Boltzmann factor $`e^{_{\mathrm{el}}/T}`$ if $`DN\zeta _{\mathrm{th}}2\theta _{\mathrm{th}}>0`$, i.e., for $`D>D_N{\displaystyle \frac{2N}{2+N}},`$ (3.21) which is equivalent to $`\zeta _\mathrm{F}>\zeta _{\mathrm{th}}`$. Note that $`D_N<2`$ for $`N<\mathrm{}`$ and $`D_{\mathrm{}}=2`$. From this observation we conclude that for $`D<D_N`$ weak disorder is irrelevant. Since on the other hand disorder will certainly become relevant if it is sufficiently strong, $`\mathrm{\Delta }>\mathrm{\Delta }_c(T)`$, one expects in this case a transition from an unpinned phase for weak disorder to a pinned phase at strong disorder (which is equivalent to a thermal depinning transition for increasing temperatures). Contrary to the result for $`\zeta _\mathrm{F}`$, which is an approximate expression for the true roughness exponent $`\zeta `$, the result for $`D_N`$ is exact. 3.1.1d Renormalization group analysis The algebraic roughness (3.6) of the manifold means that it is scale invariant on large length scales. Hence it does not have a finite correlation length and the manifold can be considered as being in a critical state with $`\zeta `$ corresponding to a critical exponent. In analogy to ordinary critical phenomena, a renormalization group (RG) analysis is suitable to describe the large-scale features of the system going beyond perturbation theory and self-consistency arguments. In $`D4`$ it is not possible to describe pinning by a finite set of parameters since $`\zeta >0`$ and a Taylor expansion of $`\mathrm{\Delta }(𝐮)`$ would yield terms the relevance of which increases with increasing order of the expansion. Therefore one has to use a functional renormalization group analysis for the present problem, which was established by \[Fisher 1986a\] (?, ?) and \[Balents and Fisher 1993\] (?) to first order in $`ϵ4D`$. On increasing length scales $`Le^l`$ the system can be described by a renormalized temperature and disorder correlator, which flow according to \[Fisher 1986a, Balents and Fisher 1993\]: $`_lT`$ $`=`$ $`\theta T,`$ (3.22a) $`_l\mathrm{\Delta }(𝐮)`$ $`=`$ $`(ϵ4\zeta )\mathrm{\Delta }(𝐮)+\zeta u_\alpha _\alpha \mathrm{\Delta }(𝐮)`$ (3.22b) $`+{\displaystyle \frac{1}{2}}_\alpha _\beta \mathrm{\Delta }(𝐮)_\alpha _\beta \mathrm{\Delta }(𝐮)_\alpha _\beta \mathrm{\Delta }(𝐮)_\alpha _\beta \mathrm{\Delta }(\mathrm{𝟎}).`$ Here we introduced a rescaled correlator $`\mathrm{\Delta }/c_\mathrm{\Delta }\mathrm{\Delta }`$ with some non-universal constant $`c_\mathrm{\Delta }`$ that depends on the short-scale cutoff and the stiffness constant $`\epsilon _{}`$. We will see that $`\theta =D2+2\zeta >0`$ such that equation (3.22a) implies that the effective temperature vanishes on large scales. The system is therefore described asymptotically by a ‘zero temperature’ fixed point. The fixed-point correlator and exponent have to be determined numerically from equation (3.22b). The actual value of the roughness exponent (‘random manifold’ value) can be represented introducing a correction factor $`\nu (D,N)`$ in the Flory expression (3.20), $$\zeta _{\mathrm{rm}}=\frac{4D}{4+\nu (D,N)N}.$$ (3.23) In all known cases this correction factor lies in the range $`0\nu (D,N)1`$. The findings of the functional RG analysis to order $`ϵ=4D`$ \[Fisher 1986a, Balents and Fisher 1993\] are equivalent to $`\nu (4,1)0.800(3)`$. It is interesting to note that the roughness can be determined exactly for special dimensions: $`\nu (D,\mathrm{})=1`$, i.e., the Flory exponent becomes exact for an infinite number of displacement components \[Mézard and Parisi 1990, Mézard and Parisi 1991\], and $`\nu (1,1)=0.5`$ corresponding to $`\zeta =\frac{2}{3}`$ \[Huse et al. 1985\]. From a more involved field-theoretical self-consistency argument \[Lässig 1998b\] (?) obtains $`\nu (1,2)=\frac{2}{5}`$ corresponding to $`\zeta =\frac{5}{8}`$ and $`\nu (1,3)=\frac{8}{21}`$ corresponding to $`\zeta =\frac{7}{12}`$. Remarkably, there is no renormalization of the elastic constant $`\epsilon _{}`$ in the Hamiltonian. The renormalized elastic constant of the manifold can be determined from its response to tilt field $`\mu _{i\alpha }^{}`$, which is coupled to the displacements through $`_\mu ={\displaystyle d^Dz\mu _{i\alpha }^{}_iu_\alpha (𝐳)}.`$ (3.24) In the absence of a pinning potential a constant tilt field leads to a manifold displacement $`u_\alpha ^{}(𝐳)=\mu _{i\alpha }^{}z_i/\epsilon _{}`$ and to a change of the free energy by an amount $`\overline{[𝝁^{\mathbf{}}]}\overline{[0]}={\displaystyle d^Dz\frac{1}{2\epsilon _{}}\mu _{i\alpha }^{}{}_{}{}^{2}}.`$ (3.25) Even in the presence of the pinning potential with a stochastic translation symmetry, this identity holds exactly. This is due to a ‘statistical symmetry’ of the pinning energy under a transformation $`𝐮(𝐳)𝐮^{}(𝐳)=𝐮(𝐳)+\delta 𝐮(𝐳)`$ for an arbitrary function $`\delta 𝐮(𝐳)`$ \[Goldschmidt and Schaub 1985, Schulz et al. 1988, Balents and Fisher 1993\], provided disorder has a vanishing correlation length in $`𝐳`$ directions, as assumed in equation (3.3). In particular, the choice $`\delta u_\alpha (𝐳)=\mu _{i\alpha }^{}z_i/\epsilon _{}`$ transforms the pinning potential $`V(𝐗+𝐮(𝐳),𝐳)V(𝐗+𝐮(𝐳)+\delta 𝐮(𝐳),𝐳)=V^{}(𝐗+𝐮(𝐳),𝐳)`$, which has the same statistical properties as the original potential. This stochastic symmetry is reflected most obviously by $`_{\mathrm{pin},n}`$ in equation (3.17), which is invariant under the transformation $`𝐮^a(𝐳)𝐮^a(𝐳)+\delta 𝐮(𝐳)`$ that is identical in all replicas. The non-renormalization of $`\epsilon _{}`$ follows directly from (3.25), since renormalized elastic constants are defined by $`{\displaystyle \frac{1}{\epsilon _{}^{\mathrm{eff}}}}={\displaystyle \frac{1}{ND}}{\displaystyle \frac{}{\mu _{i\alpha }^{}}}\overline{_iu_\alpha }={\displaystyle \frac{1}{NDL^D}}{\displaystyle \frac{^2}{\mu _{i\alpha }^{}{}_{}{}^{2}}}\overline{[𝝁^{\mathbf{}}]},`$ (3.26) the change of the free energy and the dependence of $`\overline{}`$ on $`\mu `$ is independent of disorder. Here $`L^D=d^Dz1`$ is the system size. If one starts from a disorder with a finite correlation length also in $`𝐳`$ direction, there will be a finite renormalization of the elastic constants due to fluctuations on these small length scales. Beyond this scale, the manifold behaves as for vanishing correlation length. Therefore the asymptotic behaviour on large scales (and in particular the roughness exponent) will not be affected by such correlations. The roughness exponent is, however, sensitive to the behaviour of $`\mathrm{\Delta }(𝐱)`$ on large scales, if this correlator is long ranged. Although this scenario is not of immediate interest for the present consideration of pinning by point like defects, it is realized, for example, for pinning of vortex lines by columnar defects or for magnetic domain walls with random fields. In this case (where $`N=1`$), $`\mathrm{\Delta }(x)\mathrm{\Delta }(0)|x|`$ and the roughness exponent has the ‘random-field’ value \[Villain 1984, Grinstein and Fernandez 1984, Bruinsma and Aeppli 1984\] $`\zeta _{\mathrm{rfi}}={\displaystyle \frac{4D}{3}}.`$ (3.27) Surprisingly, this exponent characterizes the manifold in the presence of a driving force at the depinning threshold, as we will discuss below. 3.1.1e Pinning vs. thermal fluctuations In section 3.1.1d, we have described the structure in the disorder-dominated regime, for the ‘zero-temperature’ fixed point. Now we reconsider the effect of thermal fluctuations in the presence of disorder. For this purpose \[Fisher and Huse 1991\] (?) separated the displacement correlation function (3.6) into two contributions, the first one due to disorder $`W_{\mathrm{pin}}(𝐳𝐳^{})\overline{𝐮(𝐳)𝐮(𝐳^{})^2},`$ (3.28) and a second one $`W_{\mathrm{th}}(𝐳𝐳^{})`$ $``$ $`W(𝐳𝐳^{})W_{\mathrm{pin}}(𝐳𝐳^{})`$ (3.29) $`=`$ $`\overline{[𝐮(𝐳)𝐮(𝐳^{})𝐮(𝐳)𝐮(𝐳^{})]^2}`$ $``$ $`{\displaystyle \frac{T}{\epsilon _{}}}|𝐳𝐳^{}|^{2D}`$ describing thermal fluctuations. In the absence of disorder, $`W_{\mathrm{pin}}(𝐳𝐳^{})=0`$. In particular $`W_{\mathrm{th}}`$ is found to be independent of disorder due to the statistical symmetry \[Schulz et al. 1988, Fisher and Huse 1991\] already mentioned above. Thus $`W_{\mathrm{th}}(𝐳)|𝐳|^{2\zeta _{\mathrm{th}}}`$ with the exponent (3.7). Consequently, $`W_{\mathrm{pin}}(𝐳)|𝐳|^{2\zeta _{\mathrm{rm}}}`$ in the low-temperature phase, where $`\zeta _{\mathrm{rm}}>\zeta _{\mathrm{th}}`$. It was argued \[Fisher and Huse 1991, Hwa and Fisher 1994a, Kinzelbach and Lässig 1995\] that the manifold would have (with probability 1) a unique ground state. Then $`W_{\mathrm{pin}}`$ essentially characterizes the ground state of the manifold \[Nattermann 1985a, Huse and Henley 1985\] and $`\zeta _{\mathrm{rm}}`$ would be the roughness exponent thereof. However, a small fraction of samples (or, small areas within a sample) will have nearly degenerate excited states with large displacements relative to the ground state. Although such excitations are rare, due to the large displacements they can dominate several disorder-averaged thermodynamic quantities \[Nattermann 1988, Hwa and Fisher 1994a\]. In particular, these rare fluctuations are responsible for the growth of the thermal width $`W_{\mathrm{th}}`$ in $`D2`$. Although $`W_{\mathrm{th}}`$ shows exactly the same behaviour as in the absence of disorder, it is important to emphasize that the distribution of thermal fluctuations is highly non-Gaussian \[Fisher and Huse 1991, Hwa and Fisher 1994a\]. For $`D>2`$ thermal fluctuations have only a finite width and the equilibrium state of the manifold globally reflects the ground state. The broken translation of the Hamiltonian are reflected by the equilibrium state of the manifold. In summary, the structure of elastic manifolds in weak disorder strongly depends on its dimensionalities and shows the following behaviour (see also figure 3): * For $`D>4`$ the manifold is flat ($`\zeta <0`$) at all temperatures provided disorder is weak. Sufficiently strong disorder will always induce roughness. * For $`2<D4`$ disorder roughens the manifold ($`\zeta =\zeta _{\mathrm{rm}}`$) at all temperatures. The manifold stays close to its ground state; the second moment of thermal displacements is finite. * For $`2DD_N=2N/(2+N)`$ disorder still roughens the manifold ($`\zeta =\zeta _{\mathrm{rm}}`$) at all temperatures. Now also the second moment of thermal displacements is infinite. * For $`D<D_N`$ and sufficiently high temperatures, the manifold is entropically driven out of the ground state and shows a structure similar to that in the absence of disorder ($`\zeta =\zeta _{\mathrm{th}}`$). However, there is also a low-temperature phase where disorder is relevant \[Cook and Derrida 1989\]. In other terms, the manifold shows a temperature-driven depinning transition at a finite temperature. For more details, the interested reader is referred to \[Halpin-Healy and Zhang 1995\] (?) and \[Lässig 1998a\] (?). #### 3.1.2 Thermodynamics So far we have described the relevance of disorder for the structure of the manifold. In particular, the roughness is qualitatively increased in the disorder dominated phases. Could such an increased roughness be taken as an unambiguous sign for glassiness of the state? To understand this problem better, we study the possibility of anomalous scaling properties of thermodynamic quantities such as energy, entropy, and free energy. Since the elastic energy and the pinning energy are local quantities, energy, entropy, and free energy scale extensively with the system size $`L`$: $`\overline{}L^D,\overline{𝒮}L^D,\overline{}L^D.`$ (3.30) Due to the randomness of pinning there are important sample-to-sample fluctuations with a scaling (for $`T>0`$) $`\overline{(\mathrm{\Delta })^2}L^D,\overline{(\mathrm{\Delta }𝒮)^2}L^D,\overline{(\mathrm{\Delta })^2}L^{2\theta },`$ (3.31) which has been confirmed numerically \[Fisher and Huse 1991\]. Here $`\theta `$ is again the energy scaling exponent (3.18). The fluctuations of $`\mathrm{\Delta }\overline{}`$ and $`\mathrm{\Delta }𝒮𝒮\overline{𝒮}`$ are normal and dominated by fluctuations on small scales: each sub-volume of the size $`L_\xi ^D`$ contributes independently. However, the fluctuations of $`\mathrm{\Delta }\overline{}`$ are anomalous and governed by large-scale contributions. Clearly, for $`T0`$ there exists a diverging length scale below which the fluctuations of $`\mathrm{\Delta }`$ scale as those of $`\mathrm{\Delta }`$. In thermodynamic equilibrium the manifold minimizes its free energy. Therefore energy and entropy are not minimized independently, their leading order cancels (at $`T>0`$), and the free energy fluctuations can be smaller than those of the energy. One expects $`\theta \frac{D}{2}`$ in all dimensions, which corresponds to $`\zeta \frac{4D}{4}`$ i.e., $`\nu (D,N)0`$ in equation (3.23). Indeed, the exponent $`\zeta _{\mathrm{IM}}=(4d)/4`$ follows from an Imry-Ma type argument \[Nattermann 1985a\] and is considered to be an upper bound for $`\zeta _{\mathrm{rm}}`$ in random-bond systems \[Fisher 1993\]. At strictly zero temperature $`=`$ and both quantities must scale in the same way, $`\mathrm{\Delta }=\mathrm{\Delta }L^\theta `$. Therefore temperature plays a crucial role, although it seems to be irrelevant in the flow equation (3.22a). This is because temperature is a dangerously irrelevant variable \[Fisher 1986b\]. In very rare samples there are excited states which are nearly degenerate with the ground state but which deviate from it by large displacements. Such rare but large fluctuations still give dominant contributions to many thermodynamic disorder-averaged thermodynamic properties, for a more detailed discussion see \[Fisher and Huse 1991\] (?) and \[Hwa and Fisher 1994a\] (?). #### 3.1.3 Susceptibility In order to find an unambiguous signature of glassy phases, \[Hwa and Fisher 1994b\] (?) proposed to examine the susceptibility of the system with respect to a tilt field. Although in this section we are dealing with directed manifolds, we establish the analogous susceptibility for these simpler systems and show how it is related to the displacement correlation functions discussed above. For simplicity we restrict the discussion of susceptibilities to $`N=1`$; the general case $`N>1`$ follows from a straightforward generalization. We couple the manifold to an inhomogeneous tilt field $`𝝁(𝐳)`$ via an additional energy contribution as in (3.24), $`_\mu ={\displaystyle d^Dz[𝝁^{\mathbf{}}+𝝁(𝐳)]\mathbf{}_{}u(𝐳)}.`$ (3.32) As in section 3.1.1d, $`𝝁^{\mathbf{}}`$ is a constant field applied to measure the tilt response of the system. The system $`^{}`$ $`=`$ $`_{\mathrm{el}}+_\mu `$ (3.33) $`=`$ $`{\displaystyle d^Dz\left\{\frac{\epsilon _{}}{2}\left(\mathbf{}_{}u\frac{1}{\epsilon _{}}[𝝁^{\mathbf{}}+𝝁]\right)^2\frac{1}{2\epsilon _{}}[𝝁^{\mathbf{}}+𝝁]^2\right\}},`$ without random potential $`V`$ but with a random tilt $`𝝁(𝐳)`$, has a ground state $`u^{}(𝐳)`$ which couples only to the longitudinal part $`𝝁_L`$ of the tilt field and which is determined by $`𝝁^{\mathbf{}}+𝝁(𝐳)=\epsilon _{}\mathbf{}_{}u^{}(𝐳)`$. For a random tilt with zero average and long-ranged correlations $`\overline{\mu _i(𝐳)\mu _j(𝐳^{})}|𝐳𝐳^{}|^\alpha ,\alpha <D,`$ (3.34) the manifold has in its ground state an exponent $`\zeta =\frac{1}{2}(2\alpha )`$ and will thus be rough if $`\alpha <2`$. Nevertheless, since this toy model is harmonic and trivial (it is bilinear in $`u`$, and $`^{}`$ is even translation invariant), it cannot be considered as glass, which motivated \[Hwa and Fisher 1994b\] (?) to search for unambiguous signatures of glassiness. They proposed to examine the sample-to-sample fluctuations of the linear response susceptibility $`\chi _{\alpha \beta }`$ of the system for spatially constant tilt $`𝝁^{\mathbf{}}`$, which can be obtained from the free energy of a particular sample by $`\chi {\displaystyle \frac{1}{DL^D}}{\displaystyle \frac{^2[𝝁^{\mathbf{}}]}{\mu _\alpha ^{}\mu _\alpha ^{}}}|_{𝝁^{\mathbf{}}=0}.`$ (3.35) Then it is interesting to consider its sample-to-sample fluctuations $`\overline{(\mathrm{\Delta }\chi )^2}\overline{\chi ^2}\overline{\chi }^2.`$ (3.36) The toy model $`^{}=_{\mathrm{el}}+_\mu `$ that has only random tilt $`𝝁`$ but no random potential, has $`\chi =1/\epsilon _{}`$ for all realizations of the random tilt because of the statistical symmetry discussed after equation (3.25). Hence the vanishing of $`\overline{(\mathrm{\Delta }\chi )^2}=0`$ reflects its trivial nature, i.e., that random tilt fields only deform the ground state, and thermal fluctuations around the ground state are similar to those in the absence of disorder. Now we come back to the manifold model in a random potential (we drop the random tilt but we still assume $`N=1`$ for simplicity) to demonstrate that $`\overline{(\mathrm{\Delta }\chi )^2}>0`$, i.e., that susceptibility fluctuations are a less ambiguous signature of glassiness than mere roughness. Due to the statistical symmetry, $`\overline{[𝝁^{\mathbf{}}]}\overline{[0]}=L^D\frac{1}{2\epsilon _{}}(𝝁^{\mathbf{}})^2`$ and $`\overline{\chi }=1/\epsilon _{}`$ exactly. We can calculate its susceptibility fluctuations perturbatively from the free energy fluctuations at $`T=0`$, $`\overline{\mathrm{\Delta }[𝝁^{\mathbf{}}]\mathrm{\Delta }[𝝁_{}^{\mathbf{}}{}_{}{}^{}]}={\displaystyle d^Dz\mathrm{\Delta }\left(\frac{1}{\epsilon _{}}(𝝁^{\mathbf{}}𝝁_{}^{\mathbf{}}{}_{}{}^{})𝐳\right)},`$ (3.37) involving the potential correlator $`\mathrm{\Delta }(𝐱)`$ evaluated for the tilted manifold in the absence of pinning. This results in $`\overline{(\mathrm{\Delta }\chi )^2}L^{4D}{\displaystyle \frac{\mathrm{\Delta }^{(4)}(0)}{\epsilon _{}^4}},`$ (3.38) where we have dropped numerical factors and $`\mathrm{\Delta }^{(4)}(0)=_u^4\mathrm{\Delta }(u)|_{u=0}`$. The dependence of this result on the system size $`L`$ shows again that $`\overline{(\mathrm{\Delta }\chi )^2}0`$ in $`D4`$. Certainly, we cannot expect this perturbative result to be quantitatively correct for large $`L`$, but qualitatively it reflects the relevance of disorder and the glassiness of the manifold. In $`D4`$ one expects $`\overline{(\mathrm{\Delta }\chi )^2}`$ to be finite. Its correct value has to be calculated within the RG scheme (see \[Hwa and Fisher 1994b\] (?)). We will not further pursue an accurate calculation of the susceptibility fluctuations here. Instead we establish its connection to the displacement fluctuations. To this end we introduce the response of the local tilt $`_\alpha u(𝐳)`$ to the field $`\mu _\beta (𝐳^{})`$ (which is now considered as an external control parameter rather than a form of disorder) $`\chi _{\alpha \beta }(𝐳,𝐳^{})`$ $``$ $`{\displaystyle \frac{\delta _\alpha u(𝐳)}{\delta \mu _\beta (𝐳^{})}}={\displaystyle \frac{\delta ^2}{\delta \mu _\beta (𝐳^{})\delta \mu _\alpha (𝐳)}}[𝝁]`$ (3.39) $`=`$ $`{\displaystyle \frac{1}{T}}[_\alpha u(𝐳)_\beta u(𝐳^{})_\alpha u(𝐳)_\beta u(𝐳^{})],`$ which is now explicitly expressed as a displacement correlation function and can be related to correlations of the order parameter $`\psi `$ exploiting $`_\alpha u(𝐳)=i_\alpha {\displaystyle \frac{\psi _k(𝐳)}{k}}|_{k=0}.`$ (3.40) The total susceptibility (3.35) discussed above can be obtained as integral of the local susceptibility (3.39), $`\chi ={\displaystyle \frac{1}{L^D}}{\displaystyle d^Dzd^Dz^{}\chi _{\alpha \alpha }(𝐳,𝐳^{})}.`$ (3.41) (Note that here is no factor $`1/D`$ in contrast to (3.35) since the displacement responds not only to the longitudinal component of $`𝝁(𝐳)`$ but to the entire $`𝝁^{\mathbf{}}`$.) For a particular sample the local susceptibility explicitly depends on two space coordinates $`𝐳`$ and $`𝐳^{}`$. We introduce a susceptibility $`\chi _{\alpha \beta }(𝐳𝐳^{})L^Dd^Dz_0\chi _{\alpha \beta }(𝐳+𝐳_0,𝐳^{}+𝐳_0)`$ averaged over the volume $`𝐳_0L^D`$, which now depends only on the coordinate difference $`𝐳𝐳^{}`$. Then $`\chi `$ is conveniently related to the Fourier transform of $`\chi (𝐳)`$ through $`\chi =\chi _{\alpha \alpha }(𝐤)|_{𝐤=0}.`$ (3.42) (For example, the toy model has $`\chi _{\alpha \beta }(𝐤)=(1/\epsilon _{})P_{\alpha \beta }^L(𝐤)`$ with the longitudinal projector $`P_{\alpha \beta }^L(𝐤)k_\alpha k_\beta /k^2`$ and hence $`\chi =1/\epsilon _{}`$.) The fact that $`\chi `$ has sample-to-sample fluctuations in a glassy phase, $`lim_L\mathrm{}\overline{(\mathrm{\Delta }\chi )^2}/\overline{\chi }^2\to ̸0`$, means that the susceptibility is not self-averaging \[Aharony and Harris 1996\]. Since $`\chi `$ is obtained from a volume average sample-to-sample fluctuations also mean that the translational average is not equivalent to the average over many samples. From the expression for the susceptibility in terms of the displacement field, equation (3.39), one recognizes that the susceptibility essentially measures fluctuations around the ground state of the system. More precisely, one can relate the susceptibility to the thermal displacement correlation (3.29) by $`_\alpha _\beta ^{}W_{\mathrm{th}}(𝐳𝐳^{})=2T\overline{\chi _{\alpha \beta }(𝐳𝐳^{})}.`$ (3.43) Thus the disorder-averaged susceptibility is related to $`W_{\mathrm{th}}`$. The fact that $`\overline{\chi }`$ is independent of disorder is related to the disorder independence of $`W_{\mathrm{th}}`$ discussed in section 3.1.1e. A signature of glassiness can therefore appear only in the sample-to-sample fluctuations of $`\chi `$. To conclude the discussion of susceptibility fluctuations as characteristic for glassiness, we wish to point out that they cannot be taken as sufficient criterion for glassiness. This can be demonstrated by a counterexample with a ‘pinning’ energy $`_{\mathrm{pin}}={\displaystyle d^Dz\frac{1}{2}[𝝂\mathbf{}_{}u(𝐳)]^2},`$ (3.44) where $`𝝂`$ is a vector of fixed length $`|𝝂|`$ but with random orientation. This type of disorder actually represents randomness in the elastic constants and the model is not a glass since it is Gaussian in the displacement. Nevertheless, it has sample-to-sample fluctuations $`\overline{(\mathrm{\Delta }\chi )^2}=[(1+𝝂^2/\epsilon _{})^{1/2}1]\overline{\chi }^2`$ (3.45) that are finite for $`𝝂^2>0`$. #### 3.1.4 Barriers The glassy nature of a system is generally related to an extremely slow dynamics that is dominated by thermally activated processes in a complex energy landscape with many meta-stable states. Glassy systems have not only a disorder-dominated ground state, but also a huge number of meta-stable states. This section is devoted to the description of the energy landscape. The dynamical behaviour is determined by ‘neighbouring’ meta-stable states, which are related to each other by the slip of a restricted part of the manifold over a certain barrier. The part of the manifold is assumed to have a certain length $`L_z`$ in $`z`$-directions, the displacement in this region is of a magnitude $`L_x`$ and the barrier height is denoted by $`U`$, see figure 4. A rigorous characterization of the statistics of barriers is very intricate. Therefore, a scaling picture \[Villain 1984, Nattermann 1985b, Huse and Henley 1985, Ioffe and Vinokur 1987\] has been put forward, where it is assumed that the statistics of barriers is essentially identical to the statistics of the free energy fluctuations. Therefore the barrier height should scale with the barrier length like $$U(L_z)U_\xi \left(\frac{L_z}{L_\xi }\right)^\psi ,$$ (3.46) where $`U_\xi U(L_\xi )\epsilon _{}\xi ^2L_\xi ^{D2}`$ is the typical height of the smallest possible barrier of the size of the Larkin length. The scaling exponent for the barrier height $`\psi `$ is assumed to coincide with the exponent $`\theta `$ from equation (3.31) that describes the free energy fluctuations: $`\psi =\theta =2\zeta +D2.`$ (3.47) This exponent is related through equation (3.18) to the roughness exponent $`\zeta `$ of the manifold, which describes the scaling relation $$L_x\xi \left(\frac{L_z}{L_\xi }\right)^\zeta $$ (3.48) between the lateral and transverse sizes of the barrier region. The basic assumption $`\psi =\theta `$ has been confirmed by analytic arguments combined with numerical simulations \[Mikheev et al. 1995, Drossel and Kardar 1995, Drossel 1996\]. However, the statistics of barriers and free energy fluctuations have turned out to possibly be not strictly identical: $`U(L_z)`$ and $`\{\overline{[\mathrm{\Delta }(L_z)]^2}\}^{1/2}`$ can differ by factors that are powers of $`\mathrm{ln}L_z`$ which do not modify the exponent relation (3.47). It is clear that not all barriers of a given size $`L_z`$ have strictly the same height but that they are distributed according to a certain distribution $`𝒫_{L_z}(U)`$. The above scaling relations are valid for the ‘optimal’ (dynamically relevant) barrier height, which is the average barrier obtained from $`𝒫_{L_z}(U)`$ with an additional weighting factor for the density of barriers of the size considered \[Vinokur et al. 1996\]. In particular in low dimensions the dynamics of the manifold is not necessarily dominated by the average barrier height, but by the largest barriers. Therefore it is also of fundamental interest to characterize the distribution $`𝒫_{L_z}(U)`$ for large $`U`$. For a string, i.e. $`(D,N)=(1,1)`$, \[Vinokur et al. 1996\] (?) found $`𝒫_{L_z}(U){\displaystyle \frac{1}{U_\xi }}{\displaystyle \frac{L_z}{L_\xi }}e^{U/U_\xi }\mathrm{exp}[(L_z/L_\xi )e^{U/U_\xi }]`$ (3.49) from a combination of extreme value statistics \[Gumbel 1958, Galambos 1978\] and a coarse graining approach. Large barriers are exponentially rare since $`𝒫_{L_z}(U)\frac{1}{U_\xi }\frac{L_z}{L_\xi }e^{U/U_\xi }`$ for $`U\mathrm{}`$. \[Gorokhov and Blatter 1999\] (?) found that the free-energy distribution obeys a similar exponential decay at large negative values of the free energies $``$. The free-energy distribution decays much faster at large positive $``$ than at large negative $``$, \[Gorokhov and Blatter 1999\], which one can expect because the manifold minimizes the free energy in equilibrium. ### 3.2 Transport properties The glassy nature of systems does not manifest itself only in equilibrium properties but, maybe even more unambiguously, in its dynamic behaviour. The existence of many meta-stable states and the thermally activated nature of transitions between these states makes the dynamics of the glass extremely slow. We describe now the transport properties of manifolds in a stationary driven state, leaving aside interesting topics such as relaxational dynamics. Vortices move dissipatively in the superconducting condensate. For small velocities the friction force is proportional to the velocity with a constant viscous drag coefficient $`\eta _0`$. For a single vortex line in a bulk superconductor \[Bardeen and Stephen 1965\] $`\eta _0{\displaystyle \frac{\mathrm{\Phi }_0H_{\mathrm{c2}}}{\rho _\mathrm{n}c^2}},`$ (3.50) where $`\rho _\mathrm{n}`$ is the normal state resistivity and $`c`$ the velocity of light. On large time scales inertial forces can be neglected in comparison to the friction force. The equation of motion is then obtained from balancing the friction force (more precisely: the force density in a $`D`$-dimensional space; for brevity forces and force densities are not strictly distinguished) with the forces arising from the interaction with other vortices, the pinning force, the driving force $`𝐅`$ and a thermal noise $`𝜻`$, $`\eta _0\dot{𝐱}={\displaystyle \frac{\delta [𝐱]}{\delta 𝐱}}+𝐅+𝜻.`$ (3.51) The noise is taken as Gaussian distributed with zero average and variance $`\zeta _\alpha (𝐳,t)\zeta _\beta (𝐳^{},t^{})=2\eta _0T\delta _{\alpha \beta }\delta (𝐳𝐳^{})\delta (tt^{})`$ (3.52) such that the equation of motion properly describes thermodynamic equilibrium according to the fluctuation-dissipation theorem in the absence of the driving force. At finite temperatures the manifold will respond to the driving force with an average velocity $`𝐯\overline{\dot{𝐱}}`$. For calculational simplicity it is advantageous to consider $`𝐯`$ as prescribed and to calculate the driving force $`𝐅=𝐅(𝐯)`$ required to maintain this velocity. In the driven case we define the displacement $`𝐮(𝐳)𝐱(𝐳)𝐗𝐯t`$ such that $`𝐮=0`$. This displacement follows the equation of motion $`\eta _0\dot{𝐮}`$ $`=`$ $`{\displaystyle \frac{\delta [𝐗+𝐯t+𝐮]}{\delta 𝐮}}+𝐅\eta _0𝐯+𝜻`$ (3.53) $`=`$ $`\epsilon _{}\mathbf{}_{}^2𝐮\mathbf{}_{}V(𝐗+𝐯t+𝐮,𝐳)+𝐅\eta _0𝐯+𝜻,`$ which serves as basis for the following analysis \[recall that $`\mathbf{}_{}(\frac{}{z_1},\mathrm{},\frac{}{z_D})`$ and $`\mathbf{}_{}(\frac{}{u_1},\mathrm{},\frac{}{u_N})`$\]. #### 3.2.1 Friction In the absence of pinning, $`V=0`$, the manifold moves with constant velocity $`𝐯=𝐅/\eta _0`$. Pinning tends to slow down the motion of the system, because the manifold has to overcome barriers and it loses more time by sliding uphill than it wins by sliding downhill. Thus, to establish a certain velocity $`𝐯`$ in the presence of pinning, one has to apply a force $`𝐅(𝐯)`$ which is larger than in the absence of pinning. In general, the velocity-force characteristic will be non-linear but one can define an effective zero-velocity friction coefficient by $`\eta _{\alpha \beta }{\displaystyle \frac{dF_\alpha (𝐯)}{dv_\beta }}|_{𝐯=\mathrm{𝟎}}.`$ (3.54) We now show that this ‘renormalized’ friction coefficient can diverge due to the presence of disorder, which is another signature of glassiness. The divergence of the effective friction coefficient is examined by treating disorder perturbatively. We make use of the linear response function of the disorder-free system $`G_{\alpha \beta }(𝐳,t)`$ $``$ $`{\displaystyle \frac{\delta u_\alpha (𝐳,t)}{\delta F_\beta (\mathrm{𝟎},0)}}={\displaystyle \frac{1}{\eta _0}}\left({\displaystyle \frac{\eta _0}{4\pi \epsilon _{}t}}\right)^{D/2}e^{\eta _0z^2/2\epsilon _{}t}\mathrm{\Theta }(t)\delta _{\alpha \beta },`$ (3.55) which describes the reaction of the displacement to a locally perturbating force. By iterating the equation of motion (3.53) up to $`O(V^2)`$, the condition $`\overline{𝐮}=0`$ yields the velocity-force characteristic and the effective friction coefficient $`\eta _{\alpha \beta }`$ $`=`$ $`\eta _0\delta _{\alpha \beta }+{\displaystyle 𝑑ttG_{\gamma \delta }(𝐳=0,t)_\alpha _\beta _\gamma _\delta \mathrm{\Delta }(\mathrm{𝟎})}.`$ (3.56) Inspecting the large-time behaviour of the response function (3.55) for $`𝐳=\mathrm{𝟎}`$, this coefficient diverges in $`D4`$ as long as $`_\alpha _\beta _\gamma _\delta \mathrm{\Delta }(\mathrm{𝟎})0`$ for certain indices. The divergence of $`\eta _{\alpha \beta }`$ resembles that of the susceptibility fluctuations (3.38): both couple to the fourth derivative of $`\mathrm{\Delta }(𝐮)`$. This divergence actually does not mean that a stationary state with non-vanishing velocity could be established only by an infinite driving force; it means that at small velocities $`𝐯(𝐅)`$ is no longer linear (Ohmic) but sub-linear. The qualitative change of the transport characteristic in $`D4`$ is consistent with roughening as a qualitative structural change. This consistency is not just a fortunate coincidence. One can actually show the relation between dynamic and static features using a fluctuation-dissipation relation between the correlation and response function of the free system \[Scheidl and Vinokur 1998b\]. Nevertheless, the toy model studied in Sec 3.1.3 has a strictly linear transport characteristic (the same as in the absence of disorder) despite its roughness. The transport characteristic is not at all affected by random bonds coupling to the displacement as in equation (3.32) because the translation symmetry $`𝐮𝐮+𝐱_0`$ is not broken. Thus the qualitative sensitivity of the transport characteristic to the nature of disorder suggests its use as a supplementary indicator for glassiness. If the dynamics of the system is glassy, the transport characteristic shows several characteristic features, cf. figure 5. For small driving forces it is important to distinguish between the cases of zero and finite temperature. At $`T=0`$ the driving force has to exceed a threshold value $`F_c`$ to set the manifold in motion with a finite average velocity, since the driving force has to overcome the average pinning force. (Strictly speaking, this is true only in the absence of quantum fluctuations, which can lead to quantum creep at $`T=0`$ and $`FF_c`$, see \[Blatter et al. 1994\] (?) and references therein.) This threshold phenomenon is called depinning. For $`T>0`$ a finite velocity is found even for $`|𝐅|<F_c`$ due to thermal activation, and the transport characteristic will be sub-linear (creep regime). At large driving forces $`|𝐅|F_c`$ the effect of pinning diminishes and the free differential mobility is reached asymptotically, $`dF_\alpha (𝐯)/dv_\beta \delta _{\alpha \beta }\eta _0`$. This latter regime is called the flow regime. The divergence of $`\eta _{\alpha \beta }`$ implies the break-down of perturbation theory for a calculation of the transport characteristic. Although it is difficult to calculate the global form of the characteristic, it is possible to achieve an analytic description for small and large velocity. #### 3.2.2 Depinning The value of the critical force $`F_c`$ where motion sets in at $`T=0`$ can be estimated as follows: On length scales below the Larkin length $`L_\xi `$ the manifold performs only displacements smaller than $`\xi `$, such that the dependence of the pinning force on the displacement can be neglected. In such a volume the local pinning forces add up to $`𝐟_\xi =_{|z|<L_\xi }d^Dz[\mathbf{}_{}V(𝐗,𝐳)]`$. The forces from different volumes are essentially uncorrelated. Thus depinning, where the driving force is balanced by pinning force, is found at \[Feigel man 1983, Bruinsma and Aeppli 1984\] $`F_cL_\xi ^D(\overline{𝐟_\xi ^2})^{1/2}(\mathrm{\Delta }^{(2)})^{1/2}L_\xi ^{D/2}\epsilon _{}\xi L_\xi ^2(\mathrm{\Delta }^{(2)})^{2/ϵ},`$ (3.57) where again $`ϵ4D`$ and the pinning force variance $`\mathrm{\Delta }^{(2)}`$ was given in equation (3.14). For fixed but weak strength of $`\mathrm{\Delta }^{(2)}`$ the critical force $`F_c`$ vanishes as $`D4`$ and it is zero for $`D>4`$. Thus for weak disorder (de)pinning occurs in $`D4`$ dimensions only, to which we restrict our analysis. The velocity is expected to increase continuously when the driving force exceeds the threshold value. This observation was made originally by \[Middleton 1992a\] (?) in the context of charge-density waves. Thus it should be possible to consider depinning as a continuous transition, at which the manifold exhibits critical fluctuations \[Nattermann et al. 1992\]. The depinning transition was studied theoretically for directed manifolds \[Nattermann et al. 1992, Narayan and Fisher 1993, Ertaş and Kardar 1996, Leschhorn et al. 1997\] and for the closely related charge-density waves \[Fisher 1985b, Narayan and Fisher 1992a\] within a renormalization group approach. \[Kardar 1998\] (?) and \[Fisher 1998\] (?) review depinning and its relevance in a much wider variety of phenomena. The phenomenology of the transition is described qualitatively by the following scenario: as long as the manifold is pinned, it will be in a rough state because of competition between elastic forces, pinning forces, and the driving force. In the absence of the driving force, the roughness is described by the exponent $`\zeta _{\mathrm{rm}}=\zeta _{\mathrm{rm}}(0)`$ of manifolds in a random potential at equilibrium, which was described in equation (3.23). A driving force polarizes the state of the manifold and turns the situation into a non-equilibrium one, where the exponent is modified to $`\zeta _{\mathrm{rm}}(F)`$. Preliminary numerical simulations \[Leaf et al. 1999\] indicate that for forces $`|𝐅|F_c`$ below the depinning transition, $`\zeta _{\mathrm{rm}}(F)`$ continuously increases with the driving force, until it reaches the value $`\zeta _{\mathrm{rm}}(F_c)=\zeta _{\mathrm{rfi}}`$ of manifolds in a random force field at equilibrium, which was described in equation (3.27). This means that the effective force acting on the manifold has qualitatively changed its character. Above depinning, where the manifold moves with a finite velocity, the force acts on large scales like an effective thermal noise (although we consider the situation at $`T=0`$, without thermal noise $`\zeta `$ in the equation of motion). Accordingly, the manifold will be flat on large scales, as long as $`D>2`$. For small velocities, the saturation of the displacement correlation function occurs on a large length scale $`\mathrm{}_z`$, which acts as a dynamical correlation length that diverges at the transition as \[Middleton 1992a, Nattermann et al. 1992, Narayan and Fisher 1993\] $`\mathrm{}_z(FF_c)^\nu `$ (3.58) with a certain exponent $`\nu `$. For this zero-temperature non-equilibrium transition, the force acts as a control parameter that plays the role of temperature in a usual finite-temperature equilibrium transition. Associated with the longitudinal scale $`\mathrm{}_z`$ there is a transverse scale $`\mathrm{}_x\mathrm{}_z^\zeta `$ (3.59) with $`\zeta `$ being the roughness exponent at depinning. As the transition is approached from above, $`FF_c`$, there is also a diverging time scale $`t_v`$, the correlation time of the pinning force acting on the manifold, $`t_v{\displaystyle \frac{\mathrm{}_x}{v}}.`$ (3.60) The dynamical exponent $`z`$ then can be defined from the scaling relation between $`\mathrm{}_z`$ and $`t_v`$, $`t_v\mathrm{}_z^z.`$ (3.61) A further scaling relation describes the continuous onset of motion, $`v(FF_c)^\beta .`$ (3.62) From a balance of the elastic and driving force in the equation of motion (making use of the non-renormalization of the elastic constant) one derives \[Nattermann et al. 1992, Narayan and Fisher 1993\] $`\nu ={\displaystyle \frac{1}{2\zeta }}.`$ (3.63) The consistency of equations (3.60), (3.61) and (3.62) requires the scaling relation $`\beta =(z\zeta )\nu .`$ (3.64) In the general case with more than one displacement component (for $`N>1`$), the velocity selects one particular direction out of the $`N`$-dimensional space and we may choose the coordinates such that $`𝐯=v\delta _{\alpha ,1}`$. Due to this selection of a direction by velocity one has to expect the displacement correlations to be uniaxially anisotropic. We subsequently distinguish $`W_\alpha (𝐳,t)=\overline{[u_\alpha (𝐳,t)u_\alpha (\mathrm{𝟎},0)]^2}`$, where $`\alpha =1`$ for the parallel component and $`\alpha =2,\mathrm{},N`$ for the perpendicular component. The anisotropy is not a minor quantitative effect, it concerns the scaling exponents as shown by \[Ertaş and Kardar 1996\] (?). The anisotropic scaling laws for the displacement components are $`W_\alpha (𝐳,t)`$ $`=`$ $`|𝐳|^{2\zeta _\alpha }g_\alpha (t/|𝐳|^{z_\alpha }),`$ (3.65) with $`g_\alpha (0)`$ finite and $`g_\alpha (y)y^{2\zeta _\alpha /z_\alpha }`$ for large arguments $`y`$. The above scaling relations (3.63) and (3.64) then hold with the identification $`\zeta \zeta _1,zz_1.`$ (3.66) The values of the exponents which can be calculated analytically within a renormalization group approach, using an expansion in small $`ϵ=4D`$ \[Nattermann et al. 1992, Ertaş and Kardar 1996\] are: $`\zeta _1`$ $`=`$ $`{\displaystyle \frac{ϵ}{3}}`$ (3.67a) $`z_1`$ $`=`$ $`2{\displaystyle \frac{2ϵ}{9}}+O(ϵ^2)`$ (3.67b) $`\zeta _\alpha `$ $`=`$ $`\zeta _1{\displaystyle \frac{D}{2}}=2+{\displaystyle \frac{5ϵ}{6}}`$ (3.67c) $`z_\alpha `$ $`=`$ $`z_1+{\displaystyle \frac{1}{\nu }}=4{\displaystyle \frac{5ϵ}{9}}+O(ϵ^2)`$ (3.67d) and all directions $`\alpha 2`$ scale identically. Equation (3.67a) means that the manifold in a random potential at depinning has the roughness of a manifold in random force field in equilibrium, cf. equation (3.27). \[Narayan and Fisher 1993\] (?) proved that equation (3.67a) is correct to all orders in $`ϵ`$. For the depinning of periodic media such as charge-density waves the same scenario holds as for directed manifolds. However, they belong to a different universality class with different values of the exponents \[Middleton 1992a, Narayan and Fisher 1992b, Narayan and Fisher 1992a\]. So far, we have discussed depinning as a phenomenon at $`T=0`$. The effect of finite temperature is that potential barriers can be overcome by thermal activation. Therefore a finite velocity is expected for every finite driving force. Fixing the force at the $`T=0`$ threshold value $`F=F_c`$, velocity increases with temperature according to a power law $`vT^{\beta /\tau }.`$ (3.68) The exponent $`\tau `$ depends on the type of the pinning potential in a non-universal way \[Fisher 1985b, Middleton 1992b\]. For the ‘ratcheted kick model’ \[Middleton 1992b\] (?) argued for $`\tau =2`$ in $`D=2,3`$. The actual value of $`\tau `$ is still controversial and subject to investigations \[Nowak and Usadel 1998, Roters et al. 1999\]. #### 3.2.3 Creep Below the depinning threshold, for $`|𝐅|F_c`$, the dynamics of the manifold is in the creep regime, where motion is possible only through thermal activation. At $`|𝐅|=F_c`$, barriers of all sizes are equally relevant and it is precisely for this reason that depinning appears as a critical phenomenon. Below depinning, barriers of a typical size are prevalent and dominate the thermally activated dynamics. In the presence of a driving force the manifold experiences an effective potential $`V_{\mathrm{eff}}(𝐱,𝐳)V(𝐱,𝐳)𝐱𝐅`$. For weak forces the manifold still has meta-stable states in this effective potential, which are separated by barriers that also depend on the driving force. We now focus on the creep regime of very small forces $`|𝐅|F_c`$, where the barriers are nearly identical to those in equilibrium and where a scaling picture for the typical, dynamically relevant barrier has been developed. The starting point of this picture is the assumption that barriers of a length $`L_z`$ and free energy fluctuations of a system of size $`L_z`$ have identical scaling as described in equation (3.46). The typical lengths $`L_z`$ and $`L_xL_z^\zeta `$ are related by equation (3.48). When the manifold overcomes such a barrier, it gains an energy $`E_FFL_xL_z^DF\xi L_\xi ^D\left({\displaystyle \frac{L_z}{L_\xi }}\right)^{D+\zeta }`$ (3.69) in the field of the driving force. Therefore, in the effective potential $`V_{\mathrm{eff}}`$ only those barriers are still effective for which $`UE_F`$. From $`U(L_F)=E_F`$ follows \[Ioffe and Vinokur 1987, Nattermann 1987, Feigelman and Vinokur 1988, Feigel’man et al. 1989, Nattermann 1990\] the size of the largest effective barrier $`L_F`$ $``$ $`L_z(F)L_\xi \left({\displaystyle \frac{F}{F_\xi }}\right)^{1/(D+\zeta \theta )}F^{1/(D+\zeta \theta )}`$ (3.70a) $`U(F)`$ $``$ $`U(L_F)U_\xi \left({\displaystyle \frac{F}{F_\xi }}\right)^\mu F^\mu `$ (3.70b) with the creep exponent $`\mu {\displaystyle \frac{\theta }{D+\zeta \theta }}={\displaystyle \frac{2\zeta +D2}{2\zeta }}.`$ (3.71) We have used the abbreviation $`F_\xi U_\xi L_\xi ^D\xi ^1`$ and the identity (3.18). The effectiveness of barriers of a given size $`L_z`$ is illustrated in figure 6. If one denotes by $`F_\mathrm{B}(L_z)U(L_z)/(L_xL_z^D)L_z^{\zeta 2}`$ the threshold force density which is required to overcome barriers of size $`L_z`$, the size of the largest effective barrier is determined alternatively by $`F=F_\mathrm{B}(L_F)`$. Since the time $`t`$ to overcome a barrier by thermal activation increases exponentially with the height $`U`$, $`t(U)e^{U/T},`$ (3.72) it is the largest effective barrier that dominates the dynamics and determines the creep velocity $`v(F)(F/\eta _0)e^{U(F)/T}(F/\eta _0)e^{(U_\xi /T)(F/F_\xi )^\mu }.`$ (3.73) The pre-exponential factor has been inserted by hand for dimensional reasons and to match the flow regime. However, in the creep regime this factor may be of more complicated form (corresponding to a logarithmic contribution to $`U(F)`$) which is not captured by this scaling argument. From equation (3.73) we see that the dynamic response will be exponentially small below depinning, since a disorder-dominated roughness with exponent $`\zeta >\zeta _{\mathrm{th}}=(2D)/2`$ implies $`\mu >0`$. The characteristic dynamic exponent $`\mu `$ depends only on the dimension and the static roughness exponent of the system. In order to provide a deeper understanding of these scaling relations several authors \[Radzihovsky 1998, Balents et al. 1998b, Chauve et al. 1998\] have recently developed a renormalization group analysis for the whole range of velocities. For $`T=0`$ their approach correctly reproduces the properties of the depinning transition and for $`T>0`$ they obtain a more precise characteristic $`v(F)`$, which confirms the exponential dependence of the characteristic (3.73) to leading order for small forces but contains also subleading orders. Experimental observations on magnetic domain walls, charge-density waves, and VLL are widely consistent with this scaling picture of barriers. The most explicit measurement of the roughness and creep exponent examining magnetic domain wall creep \[Lemerle et al. 1998\] was made only recently and is in good agreement with the theoretical results. #### 3.2.4 Flow The flow regime at large driving force $`|𝐅|F_c`$ does not usually receive much attention since it is considered trivial and is expected to be captured within perturbation theory. However, recently the flow regime has attracted some attention for driven VLL, where it turned out to be quite non-trivial and inaccessible to a perturbative analysis (see section 7) due to an intricate interplay of pinning forces and driving force. Therefore we will briefly sketch this regime also for the manifolds. Let us consider again zero temperature. There the displacements are expected to be very small for large driving forces: the manifold can respond to the pinning force only with certain (scale-dependent) relaxation times and the pinning force changes rapidly in time for large $`F`$. Therefore the effect of pinning effectively gets averaged out for $`F\mathrm{}`$. Since the disorder induced displacements are small at large $`F`$, one may again perform a perturbative analysis as a first step. According to the equation of motion (3.53) the effect of disorder is given to lowest order by a pinning force $`𝐅^{\mathrm{pin}}(𝐳,t)=\mathbf{}_{}V(𝐗+𝐯t,𝐳),`$ (3.74) which shakes the manifold as it moves over the pinning potential. Its average is zero and its variance is $`\overline{F_\alpha ^{\mathrm{pin}}(𝐳,t)F_\beta ^{\mathrm{pin}}(\mathrm{𝟎},0)}=_\alpha _\beta \mathrm{\Delta }(𝐯t)\delta (𝐳).`$ (3.75) For qualitative purposes this shaking force might be compared to an effective thermal noise. Because the velocity selects a particular space direction this noise in no longer isotropic. Effective shaking temperatures (which were introduced by \[Koshelev and Vinokur 1994\] (?) for vortex lattices) for the directions parallel and perpendicular to $`𝐯(v,\mathrm{𝟎})`$ can be identified from $`T_\alpha ^{\mathrm{sh}}`$ $`=`$ $`{\displaystyle \frac{1}{2\eta _0}}{\displaystyle 𝑑td^Dz\overline{F_\alpha ^{\mathrm{pin}}(𝐳,t)F_\alpha ^{\mathrm{pin}}(\mathrm{𝟎},0)}}`$ (3.76a) $`T_1^{\mathrm{sh}}`$ $`=`$ $`0,`$ (3.76b) $`T_\alpha ^{\mathrm{sh}}`$ $``$ $`{\displaystyle \frac{\mathrm{\Delta }_0}{2\eta _0v\xi ^{D+1}}},(\alpha 2).`$ (3.76c) These shaking temperatures are to be added to the physical temperature to give the total effective temperature, which is finite. Thus, for $`D>2`$ the manifold should be flat with exponent $`\zeta =\zeta _{\mathrm{th}}=\frac{2D}{2}`$ as in the absence of disorder. This flatness is expected to persist at any finite velocity, i.e., for all forces above $`F_c`$. However, it is interesting to observe that in the flow regime the manifold has a larger width in the direction perpendicular to $`𝐯`$ since $`T_\alpha ^{\mathrm{sh}}>T_1^{\mathrm{sh}}`$ (with $`\alpha 2`$). On the contrary, at depinning the manifold has a larger width in the direction parallel to $`𝐯`$ since $`\zeta _\alpha <\zeta _1`$, see equation (LABEL:dep.exp). This means that at a certain velocity the manifold has to reverse the aspect ratio of its widths. Apart from this interesting observation, perturbation theory gives no hints of qualitative changes of the state with changing velocity such as dynamic phase transitions. However, such transitions exist in periodic structures such as VLL (section 7). ### 3.3 Summary In this chapter we have reviewed various properties of manifolds in order as to identify a glassy state of the system. Of course, disorder always quantitatively modifies the state on small length scales. Some large-scale properties of the system can be modified qualitatively by disorder and are hence suitable for the characterization of glassiness: * Disorder reduces the positional order and it increases the roughness. Typically, $`\zeta >\zeta _{\mathrm{th}}`$ but such an increase is also possible for $`\zeta =\zeta _{\mathrm{th}}`$ if $`W`$ is increased only by logarithmic factors. * Disorder leads to large sample-to-sample fluctuations, which can be described by susceptibilities. * In a renormalization-group description of the large-scale physics, there is a disorder-dominated fixed point. Typically, this is a zero-temperature fixed point. In low dimensions ($`D2`$), this may also be a finite-temperature fixed point. * Disorder induces thermally activated sub-linear response to a driving force. We consider all these features as generic to identify the glassy nature of a system. The insufficiency of individual features can be shown by examples: An increased roughness alone is not significant as emphasized by \[Hwa and Fisher 1994b\] (?) using the bond-disorder model, which has no broken translation symmetry. Sub-linear response is a signature of broken translation symmetry but can also be achieved by ‘ordered’ potentials. Similarly, a zero-temperature fixed point can, in general, represent a low-temperature ordered phase which is not necessarily a glass. The sine-Gordon model provides an example for the last two statements. ## 4 Superconducting film in a parallel field In treating a many-vortex system we begin with the technically most simple problem of a superconducting film in a magnetic field parallel to the film plane. The mean-field phase diagram of this problem was considered by \[Abrikosov 1964\] (?), who found for $`H_{\mathrm{c1}}<H<H_{\mathrm{c2}}`$ a solution with an equidistant vortex lattice of spacing $`a`$. Here $`H_{\mathrm{c1}}=\frac{2}{\pi }\frac{\mathrm{\Phi }_0}{s^2}\mathrm{ln}\frac{s}{\xi }`$ and $`H_{\mathrm{c2}}=\frac{\sqrt{3}}{\pi }\frac{\mathrm{\Phi }_0}{\xi s}`$, where $`s`$ is the thickness of the film and we assume $`\lambda >s/\pi \xi `$ (see \[Sonin 1992\] (?) for films described by the Lawrence-Doniach model). Such a geometry can be realized experimentally and the vortex physics was observed recently on mesoscopic scales \[Bolle et al. 1999\]. In such a planar geometry vortices are not strictly confined within the film. In principle they may leave and enter the film on a side. However, such events, as well as a crossing of vortex lines within the film, always cost energy. In our theoretical analysis we will neglect such events. Then, by construction, vortex lines leave the superconducting plane only at its boundary, the vortex lattice does not exhibit topological defects like dislocations and, hence, it cannot melt unless we reach $`T_\mathrm{c}`$. Thermal fluctuations lead therefore to a quasi-long-range ordered lattice as in other 2D crystals \[Mermin and Wagner 1966\] which persists roughly up to the mean-field transition temperature $`T_{\mathrm{c0}}`$. In the following we consider the influence of frozen-in disorder on this system. It turns out that the problem can be mapped onto a magnetic model, the $`XY`$ model in a random field, for which many results are known. In addition to the phase with quasi-long-range order there is a low-temperature glassy phase, which is the result of the balance between thermal and disorder fluctuations (in this sense, the situation is different from that for bulk materials, where thermal fluctuations have a much weaker effect on the formation of the glassy state). Drift of the vortices due to an external current (perpendicular to the film plane) leads in the glassy phase to a vanishing linear resistivity. ### 4.1 Mapping on the $`XY`$ model in a random field The mapping of this model onto a two-dimensional $`XY`$ model in a random field was first considered by \[Fisher 1989\] (?) in his seminal paper on vortex glasses. Here we will follow a slightly different derivation. To formulate the model, we assume that the superconducting film is in the $`(x,z)`$ plane and that the vortex lines are directed along the $`z`$ axis with an average line spacing $`a`$. The position of the $`n`$th vortex line is denoted by $`x_n(z)=X_n+u_n(z)`$, where $`X_n=na`$ and $`u_n(z)u(X_n,z)`$ are the rest position and the displacement, respectively. The Hamiltonian can then be written in the form $`=_{\mathrm{el}}+_{\mathrm{pin}}`$, where $`_{\mathrm{el}}`$ $`=`$ $`{\displaystyle d^2r\left\{\frac{c_{11}}{2}(_xu)^2+\frac{c_{44}}{2}(_zu)^2\right\}},`$ (4.1a) $`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle d^2r\rho _u(𝐫)V(𝐫)}.`$ (4.1b) Here $`𝐫(x,z)`$, $`c_{11}`$ and $`c_{44}`$ are the compression and tilt elastic constant, $`\rho _u(𝐫)`$ the vortex line density and $`V(𝐫)`$ the random pinning potential. For the latter we assume for simplicity $`\overline{V(𝐫)}`$ $`=`$ $`0,`$ (4.2a) $`\overline{V(𝐫)V(𝐫^{})}`$ $`=`$ $`\mathrm{\Delta }(xx^{})\delta (zz^{})`$ (4.2b) and, as before, $`\mathrm{\Delta }(x)\mathrm{\Delta }_0/(\sqrt{2\pi }\xi )e^{x^2/2\xi ^2}`$, but $`\mathrm{\Delta }_0n_\mathrm{i}^{(2)}f_{\mathrm{pin}}^2\xi ^3s`$, where $`n_\mathrm{i}^{(2)}`$ and $`f_{\mathrm{pin}}`$ denote the impurity density per unit area and the individual force of a single impurity on the vortex line, respectively \[Blatter et al. 1994\]. In general, the elastic Hamiltonian will be non-local, reflecting the dispersion of the elastic constants for wave vectors $`|𝐤|>\lambda ^1`$. Since we, however, are interested in the disordered case, which is dominated by large scale fluctuations, we ignore this fact here. This approximation is justified in particular for weak disorder, a case in which all relevant length scales are much larger than $`\lambda `$. Thermal and disorder fluctuations will considerably renormalize $`c_{11}`$ and $`c_{44}`$ with respect to their mean-field values. If one starts with isolated vortex lines of displacements $`u_n`$ which have a stiffness constant $`\epsilon _{}`$ and a mutual short-ranged interaction energy $`U(a+u_{n+1}u_n)`$, then $`c_{11}=aU^{\prime \prime }(a),c_{44}=\epsilon _{}/a.`$ (4.3) On scales $`a\lambda `$ mean-field theory yields $`U(a)=U_{\mathrm{MF}}(a)e^{a/\lambda }`$. Thermal fluctuations on scales $`|\mathrm{\Delta }x|a`$ lead to a steric repulsion \[Pokrovsky and Talapov 1979\] between the vortex lines, which reads for $`\lambda 0`$ $`U_{\mathrm{th}}(a)={\displaystyle \frac{T^2}{\epsilon _{}a^2}}{\displaystyle \frac{\pi ^2}{6}},`$ (4.4) such that $`(c_{11}c_{44})^{1/2}/T=\pi /a^2`$ becomes independent of temperature. As we will see below \[cf. (4.25)\], this relation is exactly the condition for the glass transition temperature $`T_\mathrm{g}=(c_{11}c_{44})^{1/2}a^2/\pi `$. In other words, the relation (4.4) maps the system at all temperatures to the glass transition temperature $`T_\mathrm{g}`$. However, in deriving (4.4) only a hard-core repulsive interaction between vortex lines was assumed (indeed, \[Pokrovsky and Talapov 1979\] (?) mapped their problem onto non-interacting fermions). If we include an additional interaction energy of the order of $`U_c`$ per contact point of two fluctuating vortex lines, we get $`c_{11}^{\mathrm{th}}(a){\displaystyle \frac{T^2}{\epsilon _{}a^3}}\left(\pi ^2+{\displaystyle \frac{U_c}{T}}\right).`$ (4.5) This leads to $`T_\mathrm{g}(c_{11}c_{44})^{1/2}a^2/\pi =T(1+U_c/\pi ^2T)^{1/2}`$, which corresponds to $`TT_\mathrm{g}`$ for $`U_c0`$. Disorder leads also to a steric repulsion \[Kardar and Nelson 1985, Nattermann and Lipowsky 1988\] of the form (we drop here coefficients of order unity because they cannot be determined very accurately) $`U_{\mathrm{pin}}(a){\displaystyle \frac{\mathrm{\Delta }_0}{T+T_\mathrm{\Delta }}}{\displaystyle \frac{1}{a}},`$ (4.6) where $`T_\mathrm{\Delta }(\epsilon _{}\xi \mathrm{\Delta }_0)^{1/3}`$ \[Nattermann and Renz 1988\]. This gives $`{\displaystyle \frac{T_\mathrm{g}}{T}}={\displaystyle \frac{a^2}{\pi T}}\sqrt{c_{11}c_{44}}{\displaystyle \frac{T_\mathrm{\Delta }}{T}}\left({\displaystyle \frac{a}{\xi }}{\displaystyle \frac{T_\mathrm{\Delta }}{T+T_\mathrm{\Delta }}}\right)^{1/2}\sqrt{{\displaystyle \frac{a}{a_c}}}.`$ (4.7) The vortex-line distance $`a_c\xi {\displaystyle \frac{T^2(T+T_\mathrm{\Delta })}{T_\mathrm{\Delta }^3}}`$ (4.8) denotes a cross-over length scale from a thermal steric repulsion (for $`aa_c`$) to a disorder dominated steric repulsion (for $`aa_c`$). The formulas for $`U_{\mathrm{pin}}`$ and $`a_c`$ are not exact but represent crude interpolation formulas between the limiting cases $`TT_\mathrm{\Delta }`$ and $`TT_\mathrm{\Delta }`$. Finally, a contact interaction can also be included in $`U_{\mathrm{pin}}(a)`$ such that $`c_{11}^{\mathrm{pin}}(a){\displaystyle \frac{T^2}{\epsilon _{}a^2a_c}}\left(1+{\displaystyle \frac{U_c}{T}}\sqrt{{\displaystyle \frac{a_c}{a}}}\right).`$ (4.9) Here we have written $`c_{11}^{\mathrm{pin}}(a)`$ in a form similar to $`c_{11}^{\mathrm{th}}(a)`$. Comparing these two quantities, it is easy to see that $`c_{11}^{\mathrm{pin}}(a)>c_{11}^{\mathrm{th}}(a)`$ in the region $`a>a_c`$, and hence we are below $`T_\mathrm{g}`$ (provided $`U_c>0`$). One should, however, take into account that all these expressions were derived under the assumption that the bare interaction between vortex lines is short ranged, i.e., for the case $`\lambda a`$. In the opposite case the calculation is more involved and will not be considered here further. Next we consider the pinning energy (4.1b). Using the Poisson-summation formula, we may rewrite the density $`\rho _u(𝐫)`$ as (\[Nattermann 1990, Nattermann et al. 1991\]; see also appendix A) $`\rho _u(𝐫)`$ $`=`$ $`{\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}\delta (xX_nu_n(z))`$ (4.10) $`=`$ $`{\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑X\delta (XX_n)\delta (xXu_n(z))`$ $`=`$ $`{\displaystyle \frac{1}{a}}{\displaystyle \underset{m=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑Xe^{iQ_mX}\delta (xXu(X,z))`$ $`=`$ $`{\displaystyle \frac{1}{a}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑X(1+2{\displaystyle \underset{m1}{}}\mathrm{cos}(Q_mX))\delta (xXu(X,z))`$ $``$ $`{\displaystyle \frac{1}{a}}\left\{1_xu(𝐫)+2{\displaystyle \underset{m1}{}}\mathrm{cos}(Q_m[xu(𝐫)])\right\}.`$ $`Q_m2\pi m/a`$ is a reciprocal lattice vectors of the 2D line array. The pinning energy (4.1b) can hence be written as $`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle }d^2r\{{\displaystyle \frac{1}{2\pi }}_x\phi (𝐫)V(𝐫)`$ (4.11a) $`+{\displaystyle \underset{m1}{}}[h_{1m}(𝐫)\mathrm{cos}(m\phi (𝐫))+h_{2m}(𝐫)\mathrm{sin}(m\phi (𝐫))]\},`$ $`h_{1m}(𝐫)`$ $``$ $`{\displaystyle \frac{2}{a}}V(𝐫)\mathrm{cos}(Q_mx),h_{2m}(𝐫){\displaystyle \frac{2}{a}}V(𝐫)\mathrm{sin}(Q_mx),`$ (4.11b) where we introduced the phase field $`\phi 2\pi u/a`$ \[which should not be confused with the phase $`\varphi `$ of the superconducting order parameter; see equation (4.38) below\]. The disorder fields $`h_{\alpha m}`$ ($`\alpha ,\beta =1,2`$) are Gaussian distributed with $`\overline{h_{\alpha m}(𝐫)}`$ $`=`$ $`0,`$ (4.12a) $`\overline{h_{\alpha m}(𝐫)h_{\beta n}(𝐫^{})}`$ $``$ $`2{\displaystyle \frac{\mathrm{\Delta }_0}{a^2}}f_{\alpha m,\beta n}(Q_mx,Q_nx^{})\delta _\xi (𝐫𝐫^{}).`$ (4.12b) On scales $`\mathrm{\Delta }xa`$, $`f_{\alpha m,\beta n}(Q_mx,Q_nx^{})\delta _{\alpha \beta }\delta _{mn}`$ since the rapidly fluctuating contributions with $`mn`$ or $`\alpha \beta `$ average to zero. On these scales, $`_{\mathrm{pin}}`$ can also be written in the form $`_{\mathrm{pin}}{\displaystyle d^2r\left\{\frac{1}{2\pi }_x\phi (𝐫)V(𝐫)+\underset{m1}{}\frac{2}{a}\mathrm{\Delta }_0^{1/2}\mathrm{cos}(m\phi (𝐫)\alpha _m(𝐫))\right\}}.`$ (4.13) The random phases $`\alpha _m(𝐫)`$ obey $`\overline{e^{i\alpha _m(𝐫)}}`$ $`=`$ $`0,`$ (4.14a) $`\overline{e^{i[\alpha _m(𝐫)\alpha _n(𝐫^{})]}}`$ $`=`$ $`\delta _{mn}\delta _\xi (𝐫𝐫^{}).`$ (4.14b) The effect of fluctuations of individual vortex lines on scales $`\mathrm{\Delta }xa`$ was discussed briefly in the first part of this section. Since we study the film in a parallel field mainly because of the possibility of obtaining explicit results for the vortex-glass state in a case which we understand to a large extent (and not so much because of its experimental relevance), we will ignore collective fluctuation effects on small and intermediate scales but concentrate directly on the large-scale behaviour where typical phase fluctuations are of the order $`2\pi `$. On these scales terms with $`m>1`$ in (4.13) can be ignored since they are less relevant than that with $`m=1`$. It is convenient to rewrite (4.1a) as $`_{\mathrm{el}}`$ $`=`$ $`{\displaystyle \frac{J}{2}}{\displaystyle d^2r(\mathbf{}\phi )^2}`$ (4.15) by rescaling the $`z`$ coordinate and introducing the stiffness constant $`J`$ according to $`\stackrel{~}{z}`$ $``$ $`\left({\displaystyle \frac{c_{11}}{c_{44}}}\right)^{1/2}z,`$ (4.16a) $`J`$ $``$ $`\left({\displaystyle \frac{a}{2\pi }}\right)^2\sqrt{c_{11}c_{44}}.`$ (4.16b) This rescaling leaves $`_{\mathrm{pin}}`$ form-invariant apart from a multiplication with a factor $`(c_{44}/c_{11})^{1/4}`$. Thus $`_{\mathrm{el}}+_{\mathrm{pin}}`$ now equals the Hamiltonian of a random-field (RF) $`XY`$ model without vortices, which was considered in the context of magnetic models by a number of authors \[Cardy and Ostlund 1982, Goldschmidt and Houghton 1982, Goldschmidt and Schaub 1984, Villain and Fernandez 1984\]. After rescaling (4.16a), $`_{\mathrm{pin}}`$ can be written as $`_{\mathrm{pin}}={\displaystyle d^2r\left\{\stackrel{~}{V}(\phi (𝐫),𝐫)𝝁(𝐫)\mathbf{}\phi (𝐫)\right\}}`$ (4.17) with $`\overline{𝝁}`$ $`=`$ $`\overline{\stackrel{~}{V}}=0,`$ (4.18a) $`\overline{\mu _i(𝐫)\mu _j(𝐫^{})}`$ $`=`$ $`J^2\sigma \delta _{ij}\delta _\xi (𝐫𝐫^{}),`$ (4.18b) $`\overline{\stackrel{~}{V}(\phi ,𝐫)\stackrel{~}{V}(\phi ^{},𝐫^{})}`$ $`=`$ $`h^2\mathrm{cos}(\phi \phi ^{})\delta _\xi (𝐫𝐫^{}),`$ (4.18c) where we introduced the variances $`\sigma `$ and $`h^2`$. Their bare (unrenormalized) values are given by $`h_0^2`$ $``$ $`2{\displaystyle \frac{\mathrm{\Delta }_0}{a^2}}\left({\displaystyle \frac{c_{44}}{c_{11}}}\right)^{1/2},`$ (4.19a) $`\sigma _0`$ $``$ $`{\displaystyle \frac{h_0^2a^2}{8\pi J^2}}.`$ (4.19b) This relation between $`\sigma `$ and $`h^2`$ in (LABEL:def.h.sig) holds only for the unrenormalized quantities. However, under renormalization these quantities flow independently. Although (4.13) originally contained only a linear gradient term in the $`x`$ direction, we rewrote it here in an isotropic form, anticipating the generation of $`_z\phi `$ terms under the renormalization-group transformation. Before we come to the renormalization-group calculation for the model (4.17), we will apply a Flory-type analysis to it. To be more general, we will present this analysis in $`d`$ dimensions. One complication with respect to the problem considered in section 3.1.1c is the fact that the correlator of the pinning Hamiltonian now contains the oscillatory term (4.18c). For weak disorder $`haJ`$, a given configuration of the random phase $`\alpha _1(𝐫)\alpha (𝐫)`$ implies a certain ground-state configuration $`\phi _0(𝐫)=\phi _0(𝐫,\{\alpha \})`$ which depends on the value of the random phase everywhere in the system. Going through the different configurations of $`\{\alpha \}`$ will generate a distribution of ground state configurations $`\{\phi _0\}`$ which we assume to be Gaussian distributed. Since $`haJ`$, the correlation of the value of $`\alpha (𝐫)`$ and $`\phi _0(𝐫)`$ at the same position $`𝐫`$ will be weak. In averaging over the distribution of $`\alpha (𝐫)`$ we can therefore neglect the local correlations between $`\alpha (𝐫)`$ and $`\phi _0(𝐫)`$ in a first approximation. The typical energy gain from the pinning term then becomes $`E_{\mathrm{pin}}\mathrm{max}\{J\sqrt{\sigma }L^{(d2)/2}(\overline{\phi ^2})^{1/2},hL^{d/2}e^{\overline{\phi ^2}/2}\}.`$ (4.20) Depending on the dimension $`d`$, the leading contribution to $`E_{\mathrm{pin}}`$ can be the random-tilt or the random-field contribution. The dependence of $`\overline{\phi ^2}`$ on the system size $`L`$ follows from equating the absolute value of $`E_{\mathrm{pin}}`$ to the averaged elastic energy $`E_{\mathrm{el}}JL^{d2}\overline{\phi ^2}.`$ (4.21) For $`2<d<4`$ one finds that the random-field contribution dominates in $`E_{\mathrm{pin}}`$, which results in $`\overline{\phi ^2}(4d)\mathrm{ln}{\displaystyle \frac{L}{L_\xi }}2\mathrm{ln}[\overline{\phi ^2}],`$ (4.22) where we introduced the Larkin length $`L_\xi (J/h)^{2/(4d)}`$. Thus, the Flory argument gives at $`T=0`$ a logarithmic increase of $`\overline{\phi ^2}`$ in all $`2<d<4`$. We will see in section 6 that a renormalization-group calculation confirms essentially this result, apart from a (small) modification of the coefficient of the logarithm. For $`d<2`$ on the other hand, $`E_{\mathrm{pin}}`$ is dominated by the random-tilt contribution and the balance of $`E_{\mathrm{pin}}`$ and $`E_{\mathrm{el}}`$ at $`T=0`$ results in $`\overline{\phi ^2}\sigma L^{2d}`$ in agreement with renormalization-group calculations \[Villain and Fernandez 1984\]. We will show below that in $`d=2`$ dimensions $`\sigma `$ is renormalized to large values such that $`E_{\mathrm{pin}}`$ and $`\overline{\phi ^2}`$ are dominated by the random-tilt contribution. ### 4.2 Renormalization After replication, we get from (4.17) and (LABEL:mu.corr) $`_n={\displaystyle \underset{ab}{}}{\displaystyle d^2r\left\{\frac{1}{2}J(\delta ^{ab}\frac{J\sigma }{T})\mathbf{}\phi ^a\mathbf{}\phi ^b\frac{h^2}{2T}\mathrm{cos}(\phi ^a\phi ^b)\right\}}.`$ (4.23) The RG flow equations of this model, first found by \[Cardy and Ostlund 1982\] (?), are then $`{\displaystyle \frac{dJ}{dl}}`$ $`=`$ $`0,`$ (4.24a) $`{\displaystyle \frac{d\sigma }{dl}}`$ $`=`$ $`c_1{\displaystyle \frac{h^4}{T^2J^2}},`$ (4.24b) $`{\displaystyle \frac{dh}{dl}}`$ $`=`$ $`\left(1{\displaystyle \frac{T}{T_\mathrm{g}}}\right)hc_2{\displaystyle \frac{h^3}{T^2}},`$ (4.24c) with the glass transition temperature $`T_\mathrm{g}=4\pi J.`$ (4.25) The integration of (4.24c) yields $`h^2(L)=h_0^2\left({\displaystyle \frac{L}{a}}\right)^{2\tau _\mathrm{g}}\left\{1+{\displaystyle \frac{h_0^2}{h_{}^{}{}_{}{}^{2}}}\left[\left({\displaystyle \frac{L}{a}}\right)^{2\tau _\mathrm{g}}1\right]\right\}^1`$ (4.26) with $`l=\mathrm{ln}(L/a)`$, the fixed-point value $`h_{}^{}{}_{}{}^{2}=T^2\tau _\mathrm{g}/c_2`$ and the reduced temperature $`\tau _\mathrm{g}1{\displaystyle \frac{T}{T_\mathrm{g}}}.`$ (4.27) It can be shown quite generally that the flow of $`\sigma `$ does not feed back into that of $`h`$ \[Hwa and Fisher 1994b\]. The constants $`c_1`$ and $`c_2`$ are cut-off dependent, but their ratio $`{\displaystyle \frac{c_1}{c_2^2}}={\displaystyle \frac{1}{4\pi }}+O(\tau _\mathrm{g})`$ (4.28) is universal. The whole calculation is valid only for $`\tau _\mathrm{g}1`$. Considering equivalent models, equivalent RG equations were found by \[Goldschmidt and Houghton 1982\] (?), \[Rubinstein et al. 1983\] (?), \[Goldschmidt and Schaub 1984\] (?), and \[Paczuski and Kardar 1991\] (?). The parameter flow is shown schematically in figure 7. The quantity $`h^2`$ approaches a fixed point $`h_{}^{}{}_{}{}^{2}T_\mathrm{g}^2\tau _\mathrm{g}/c_2`$. Integrating (4.24b) for $`hh^{}`$, we find on the length scale $`L=ae^l`$ $`\sigma (L)4\pi \tau _\mathrm{g}^2\mathrm{ln}(L/a).`$ (4.29) If we go back to the renormalized but unrescaled quantities – these are the physical quantities we would measure in an experiment on this scale – we obtain for the effective parameters on scale $`L`$ $`h_{\mathrm{eff}}`$ $``$ $`h(L){\displaystyle \frac{a}{L}},`$ (4.30a) $`\sigma _{\mathrm{eff}}`$ $`=`$ $`\sigma (L).`$ (4.30b) Thus, while the effective random-field strength vanishes as $`L^1`$, the variance of the coefficient of the linear gradient term grows logarithmically with length scale $`L`$. As a side remark we mention here that the effective value of the coefficient of the non-linear term in $`_n`$ scales as $`L^2`$ in agreement with the general result for $`\mathrm{\Delta }_{\mathrm{eff}}h_{\mathrm{eff}}^2L^{d4}`$, cf. equation (6.28) below. It is interesting that \[Villain and Fernandez 1984\] (?) considered this problem at $`T=0`$ in a complementary study and found the following set of RG flow equations (we rewrite here their discrete recursion relations in a differential form): $`{\displaystyle \frac{dJ}{dl}}`$ $`=`$ $`0,`$ (4.31a) $`{\displaystyle \frac{d\sigma }{dl}}`$ $`=`$ $`\stackrel{~}{c}_1{\displaystyle \frac{h^2}{J^2+\stackrel{~}{c}_2h^2}},`$ (4.31b) $`{\displaystyle \frac{dh}{dl}}`$ $`=`$ $`{\displaystyle \frac{hJ^2}{J^2+\stackrel{~}{c}_2h^2}},`$ (4.31c) with some numerical constants $`\stackrel{~}{c}_1`$, $`\stackrel{~}{c}_2`$. $`h`$ flows now to infinity, $`h(L)\sqrt{2/\stackrel{~}{c}_2}J\mathrm{ln}L`$, such that $`d\sigma /dl\stackrel{~}{c}_1/\stackrel{~}{c}_2`$. Thus both calculations, for $`TT_\mathrm{g}`$ and for $`T=0`$, give a logarithmic increase of $`\sigma _{\mathrm{eff}}(L)`$ and a vanishing $`h_{\mathrm{eff}}(L)`$ for $`L\mathrm{}`$. This gives further credibility to the existence of a unique low-temperature phase. A central role for the characterization of this phase plays the displacement correlation function \[Goldschmidt and Houghton 1982\] $`W(𝐫)\overline{[\phi (𝐫)\phi (\mathrm{𝟎})]^2}={\displaystyle \frac{T}{\pi J}}\mathrm{ln}{\displaystyle \frac{r}{a}}+2\tau _\mathrm{g}^2\mathrm{ln}^2{\displaystyle \frac{r}{a}}.`$ (4.32) Qualitatively, the same result was found by \[Villain and Fernandez 1984\] (?) at $`T=0`$. From (4.32) it is possible to calculate the correlation function of the order parameter for translational order $`\psi _k(𝐫)=e^{iku(𝐫)}`$ \[cf. equation (3.8)\]. For $`k=Q_m=2\pi m/a`$ one has $`ku=m\phi `$ and hence \[Goldschmidt and Houghton 1982\] $`S(Q_m,𝐫)`$ $`=`$ $`\overline{e^{iQ_m[u(𝐫)u(\mathrm{𝟎})]}}=\overline{e^{im[\phi (𝐫)\phi (\mathrm{𝟎})]}}`$ (4.33) $``$ $`e^{\frac{m^2}{2}\overline{[\phi (𝐫)\phi (\mathrm{𝟎})]^2}}r^{m^2\eta (r)},`$ where $`\eta (r)={\displaystyle \frac{T}{2\pi J}}+2\tau _\mathrm{g}^2\mathrm{ln}{\displaystyle \frac{r}{a}},T<T_\mathrm{g}.`$ (4.34) Thus, in the glassy phase, at $`T<T_\mathrm{g}`$, the correlation function $`S(Q,𝐫)`$ decays slightly faster than with a power law. However, the $`\mathrm{ln}^2r`$-behaviour dominates $`W(𝐫)`$ only on scales $`|𝐫|L_{\tau _\mathrm{g}}ae^{2/\tau _\mathrm{g}^2}`$ which is large within the range of validity $`\tau _\mathrm{g}1`$ of the RG equations (LABEL:CO.flow). The structure factor $`\widehat{S}(Q+q_x,q_z){\displaystyle \frac{1}{a}}{\displaystyle d^2re^{i(q_xx+q_zz)}S(Q,𝐫)}`$ (4.35) behaves for small $`𝐪=(q_x,\sqrt{c_{44}/c_{11}}q_z)`$ near the first reciprocal lattice vector $`Q=\pm 2\pi /a`$, therefore, as $`\widehat{S}(Q+q_x,q_z)q^{2\tau _\mathrm{g}[1\tau _\mathrm{g}\mathrm{ln}(aq)]}.`$ (4.36) Contrary to the pure system, where quasi-long-range order is accompanied by algebraic Bragg peaks, these are smeared out here for $`qL_{\tau _\mathrm{g}}^1`$. For $`T>T_\mathrm{g}`$, the term $`\mathrm{ln}r`$ in (4.34) is absent and the power law decay of correlations is regained. For completeness, and since the pair correlation function $`S(Q,𝐫)`$ vanishes faster than a power law, we also consider the positional glass correlation function $`S_{\mathrm{PG}}(Q,𝐫)`$. cf. equation (2.33). This was also calculated already by \[Goldschmidt and Houghton 1982\] (?), who found $$S_{\mathrm{PG}}(Q,𝐫)|𝐫|^{4T/T_\mathrm{g}}$$ (4.37) for $`|\tau _\mathrm{g}|1`$ on both sides of the transition. Note that to the given accuracy of order $`\tau _\mathrm{g}`$ this glass correlation function decays as a power law in both phases. In the high-temperature phase, where disorder is essentially irrelevant, the lack of a mechanism for melting in the model – in particular the absence of dislocations – prevents $`S_{\mathrm{PG}}`$ from vanishing. In the low-temperature glassy phase, on the other hand, thermal fluctuations are still too strong to allow for true long-range order of the glass correlation function in this two-dimensional system. Note, however, that the relative magnitude of the correlation functions is qualitatively different in both phases, since $`S_{\mathrm{PG}}/S^2const.`$ for $`T>T_\mathrm{g}`$ but $`S_{\mathrm{PG}}/S^2\mathrm{}`$ for $`T<T_\mathrm{g}`$ and $`|𝐫|\mathrm{}`$. We finally remark here that for this simple model, which has no shear modes, also the phase-coherent vortex glass correlation $`C_{\mathrm{VG}}(𝐫)`$ (2.34) shows an algebraic decay. Indeed, since in this layered geometry the phase $`\varphi `$ of the superconducting order parameter changes by $`\pi `$ at a vortex line, the latter is related to the displacement by $$\varphi (x,z)\frac{\pi }{a}\left(xu(x,z)\right)=\frac{\pi x}{a}\frac{1}{2}\phi (x,z).$$ (4.38) With (2.34) and (4.37) we find therefore $$C_{\mathrm{VG}}(𝐫)|𝐫|^{T/T_\mathrm{g}},$$ (4.39) i.e., the system shows quasi-long-range vortex-glass order. Our result (4.32) for the glassy phase is in contradiction with a number of results by other authors using Bethe Ansatz \[Tsvelik 1992, Balents and Kardar 1993\], a variational treatment with one-step replica-symmetry breaking \[Korshunov 1993, Giamarchi and Le Doussal 1994\], and variational methods without replicas \[Orland and Shapir 1995\]. In these studies $`W(𝐫)\mathrm{ln}r`$ was also found for the low-temperature phase, but with a temperature independent coefficient for low $`T`$. We believe that these results are incorrect and in particular that they demonstrate the flaws of the Gaussian variational method, which gives only Flory-like results for the correlation function. Indeed, \[Bauer and Bernard 1996\] (?) showed for an $`N`$-component version of our model (4.23), that the coefficient of the $`\mathrm{ln}^2r`$-term in $`W(𝐫)`$ vanishes as $`1/N^3`$ for large $`N`$. This explains the absence of the $`\mathrm{ln}^2r`$-term in the variational calculations which give exact results only in the limit $`N\mathrm{}`$. Numerical studies give a controversial picture: while in the investigations of \[Batrouni and Hwa 1994\] (?) and \[Cule and Shapir 1995\] (?) only a $`\mathrm{ln}r`$-behaviour of $`W(𝐫)`$ was found, later studies \[Lancaster and Ruiz-Lorenzo 1995, Marinari et al. 1995, Zeng et al. 1996, Rieger and Blasum 1997\] were able to detect a $`\mathrm{ln}^2r`$-behaviour. However, the coefficient of the $`\mathrm{ln}^2r`$ term found in the finite-temperature simulations is much smaller than the RG prediction (4.32), even if one takes into account that the true coefficient $`2\tau _\mathrm{g}^2`$ is smaller by a factor $`\frac{1}{4}`$ than assumed originally. The difficulty in observing the latter behaviour may partially be explained by a large crossover length. More recently, \[Zeng et al. 1999b\] (?) were able to confirm in their simulation both the $`\mathrm{ln}^2r`$ dependence and the $`\tau _\mathrm{g}^2`$ prefactor of the correlation $`W(𝐫)`$. The faster decay of correlations of the order parameter for translational order in the low-temperature phase clearly indicates the influence of disorder. However, the resulting phase is not necessarily a glassy phase. Indeed, as was remarked by \[Hwa and Fisher 1994b\] (?) and mentioned in connection with our discussion of manifolds, a harmonic toy model (4.15), (4.17) with $`\stackrel{~}{V}(\phi ,𝐫)0`$ but \[compare (3.34)\] $`\overline{\mu _i(𝐫)\mu _j(𝐫^{})}=\delta _{ij}J^2\sigma f(𝐫𝐫^{}),f(𝐫)|𝐫|^\alpha ,\alpha <d`$ (4.40) already results in $`W(𝐫)=\overline{[\phi (𝐫)\phi (\mathrm{𝟎})]^2}\sigma |𝐫|^{2\alpha }.`$ (4.41) Although the roughness exponent $`\zeta =\frac{1}{2}(2\alpha )`$ can be larger than zero in $`d=2`$ dimensions, and hence $`S(Q,𝐫)`$ would decay like a stretched exponential function, the system is certainly not glassy since it has indeed a continuum of ground states which differ by the value of a constant $`\phi _1`$ but have the same energy. Thus, there is no pinning of the vortex-line lattice in this toy model. We conclude from this exercise that a stronger decay of the correlation function $`S(Q,𝐫)`$ than that resulting from thermal fluctuations is a necessary but not sufficient requirement for the existence of a glassy phase. It is clear that the existence of many meta-stable states and, hence, the anharmonicity of the model is decisive. ### 4.3 Susceptibility As a better signature of glassiness \[Hwa and Fisher 1994b\] (?) proposed to look at the response to a change $`\delta 𝐇=(\delta H_x,\delta H_z)`$ in the applied field. This change leads to a new term $`(\mathrm{\Phi }_0/8\pi ^2)d^2r\delta 𝐇\mathbf{}\phi `$ in the Hamiltonian. One can now consider the susceptibility $`\chi _{ij}{\displaystyle \frac{}{h_j}}_i\phi .`$ (4.42) In isotropic systems $`\chi _{ij}=\chi \delta _{ij}`$ and the magnetic permeability is found as $`\mu _{\mathrm{mag}}=\chi \mathrm{\Phi }_0^2/(16\pi ^3).`$ (4.43) It was then shown that the sample-to-sample fluctuations of $`\chi `$ or $`\mu _{\mathrm{mag}}`$ in the low-temperature phase fulfill the relation $`\overline{(\mathrm{\Delta }\chi )^2}\overline{(\chi \overline{\chi })^2}=C\tau _\mathrm{g}\overline{\chi }^2,(T<T_\mathrm{g}),`$ (4.44) where $`\overline{\chi }=1/J`$ is independent of the disorder as a result of an underlying statistical tilt symmetry \[Schulz et al. 1988\]. $`C`$ is a universal boundary- and geometry-dependent coefficient and $`\tau _\mathrm{g}c_2(h^{}/T_\mathrm{g})^2`$ \[cf. equation (4.27)\] is a measure of the effective non-linearity of the model on large length scales. On the contrary, $`\overline{(\mathrm{\Delta }\chi )^2}0`$ for $`T>T_\mathrm{g}`$ and for large systems. The non-vanishing sample-to-sample fluctuations of the susceptibility for $`T<T_\mathrm{g}`$, which may be tested already at a simple sample by rotating the direction of the magnetic field, are therefore a better way to define glassiness. Although early numerical simulations by \[Batrouni and Hwa 1994\] (?) were not able to confirm prediction (4.44), more recent simulations by \[Zeng et al. 1999b\] (?) are in agreement with (4.44). For completeness we mention in addition a study by \[Sudbø 1993\] (?), in which he considers the response $`D\chi _{xx}`$ of the vortex-line array to a tilt of the magnetic field. By mapping the system of hard-core repelling vortex lines onto non-interacting 1D fermions in a time dependent random potential, he identifies a possible glassy phase of vortex lines with the localization of fermions. For electrons in a time independent random potential one expects a vanishing average charge stiffness $`\overline{D}`$ in the localized phase, in contrast to his Bethe-Ansatz calculation, which gives $`\overline{D}0`$. However, as mentioned before, $`\overline{D}0`$ is the consequence of an underlying statistical tilt symmetry (which exists only for a fermion potential which is random both in space and in time direction) and not a signature of the absence of a glassy phase \[Hwa and Nattermann 1993\]. ### 4.4 Dynamics The physically most convincing way to demonstrate glassy properties of a system is to consider the dynamics. Before we begin this topic, we will make a small digression to discuss the appearance of electrical resistivity from the motion of vortex lines. Under the influence of an external current density $`𝐣`$ a Lorentz force with density $`𝐅={\displaystyle \frac{1}{c}}𝐣𝐁`$ (4.45) acts on the vortex-line array. $`𝐁`$ denotes its magnetic induction and $`c`$ the velocity of light. The force density leads under dissipative conditions to a steady-state motion of vortex lines of velocity $`𝐯=𝐯(𝐅)`$, which generates an electric field $`𝐄=𝐁𝐯/c`$. If $`𝐯`$ and $`𝐅`$ are parallel to each other, $`𝐄`$ and $`𝐣`$ are as well. The resistivity of the vortex-line array follows, therefore, from $`\rho (j)={\displaystyle \frac{dE}{dj}}={\displaystyle \frac{B^2}{c^2}}{\displaystyle \frac{dv}{dF}}.`$ (4.46) A simple situation exists if the relation $`𝐯(𝐅)=𝐅/\eta `$ is linear with a friction coefficient $`\eta `$, for which $`\rho =B^2/(c^2\eta )`$ holds. An example for a linear relation is the Bardeen-Stephen flux flow where $`\eta =BH_{\mathrm{c2}}/(c^2\rho _\mathrm{n})`$ with the normal resistivity $`\rho _\mathrm{n}`$ \[Bardeen and Stephen 1965\]. We consider now the influence of disorder in the 2D case with the magnetic field parallel to the film plane. The equation of motion of the vortex-line array under the influence of a driving force density $`F`$ in $`x`$ direction (the external current is assumed to be perpendicular to the film) reads $`\eta _0\dot{\phi }`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta \phi }}+f+\zeta ,`$ (4.47a) $`\zeta (𝐫,t)\zeta (𝐫^{},t^{})`$ $`=`$ $`2\eta _0T\delta (𝐫𝐫^{})\delta (tt^{}),`$ (4.47b) with $`fF`$ and $`\eta _0BH_{\mathrm{c2}}/\rho _\mathrm{n}`$. From a dynamical RG one finds \[Goldschmidt and Schaub 1985, Tsai and Shapir 1992, Tsai and Shapir 1994\] a renormalization of the friction coefficient according to $`{\displaystyle \frac{d\eta }{dl}}=4c_2\sqrt{c_\mathrm{g}}{\displaystyle \frac{h^2}{T_\mathrm{g}^2}}\eta ,`$ (4.48) where $`c_\mathrm{g}e^{2\gamma _{\mathrm{Eu}}}/4`$ and $`\gamma _{\mathrm{Eu}}`$ is Euler’s constant, i.e., $`2\sqrt{c_\mathrm{g}}1.78`$. $`c_2`$ was introduced in equation (LABEL:CO.flowc). The integration of (4.48) gives for the effective friction constant on the scale $`L=ae^l`$ $`\eta (L)\eta _0\left[1+{\displaystyle \frac{c_2}{\tau _\mathrm{g}}}\left({\displaystyle \frac{h_0}{T_\mathrm{g}}}\right)^2\left((L/a)^{2\tau _\mathrm{g}}1\right)\right]^{2\sqrt{c_\mathrm{g}}}L^{z2}`$ (4.49) for $`T<T_\mathrm{g}`$, where we used (LABEL:CO.flowc). Because of the absence of a trivial rescaling in (4.48), $`\eta (L)`$ coincides with the effective friction coefficient $`\eta _{\mathrm{eff}}(L)`$. The relation $`\eta (L)L^{z2}`$ defines the dynamical critical exponent $`z`$ with $`z=\{\begin{array}{cc}2+4\sqrt{c_\mathrm{g}}\tau _\mathrm{g}\hfill & \text{ for }T<T_\mathrm{g}\hfill \\ 2\hfill & \text{ for }TT_\mathrm{g}.\hfill \end{array}`$ (4.52) from (4.49). For $`T>T_\mathrm{g}`$, $`\tau _\mathrm{g}<0`$ and the asymptotic ($`L\mathrm{}`$) value $`\eta _{\mathrm{}}`$ of $`\eta `$ is $`\eta _{\mathrm{}}(\tau _\mathrm{g})\eta _0\left(1+c_2{\displaystyle \frac{h_0^2}{T_\mathrm{g}^2}}|\tau _\mathrm{g}|^1\right)^{2\sqrt{c_\mathrm{g}}}|\tau _\mathrm{g}|^{2\sqrt{c_\mathrm{g}}}.`$ (4.53) From this one finds for the linear resistivity close to $`T_\mathrm{g}`$ $`\rho _{\mathrm{}}\rho _\mathrm{n}{\displaystyle \frac{B}{H_{\mathrm{c2}}}}\left(1+c_2{\displaystyle \frac{h_0^2}{T_\mathrm{g}^2}}|\tau _\mathrm{g}|^1\right)^{2\sqrt{c_\mathrm{g}}}|\tau _\mathrm{g}|^{2\sqrt{c_\mathrm{g}}}.`$ (4.54) In other words, we obtain for the depinned solid phase a linear (Ohmic) resistivity which vanishes at $`T_\mathrm{g}`$ as a power law in $`|\tau _\mathrm{g}|`$ with the exponent $`2\sqrt{c_\mathrm{g}}`$. For $`T<T_\mathrm{g}`$, $`\tau _\mathrm{g}>0`$ and $`\eta (L)`$ increases with increasing length scale until $`LL_{\mathrm{max}}a(J/F)^{1/2}`$ is reached where the RG flow is stopped \[Tsai and Shapir 1992\]. The response is therefore non-linear with ($`FJ`$) $`\eta _{\mathrm{eff}}(\tau _\mathrm{g}>0,F)\eta _0\left[1+{\displaystyle \frac{c_2}{\tau _\mathrm{g}}}\left({\displaystyle \frac{h_0}{T_\mathrm{g}}}\right)^2\left((J/F)^{\tau _\mathrm{g}}1\right)\right]^{2\sqrt{c_\mathrm{g}}},`$ (4.55) which gives a non-linear resistivity $`\rho _{\mathrm{eff}}(j)\rho _n{\displaystyle \frac{B}{H_{\mathrm{c2}}}}\left[1+{\displaystyle \frac{c_3}{\tau _\mathrm{g}}}\left(\left({\displaystyle \frac{j_c}{j}}\right)^{\tau _\mathrm{g}}1\right)\right]^{2\sqrt{c_\mathrm{g}}}j^{(z2)/2},`$ (4.56) where $`j_c`$ has the meaning of a zero-temperature critical current density, $`c_3=O(1)`$ is a constant and $`jj_c`$. Apparently, the linear resistivity $`\rho _{\mathrm{eff}}(j0)`$ vanishes, i.e., the system in the glassy phase is a true superconductor. For $`\tau _\mathrm{g}0`$ we obtain in particular at the vortex-glass transition $`\rho _{\mathrm{eff}}(j)\rho _\mathrm{n}{\displaystyle \frac{B}{H_{\mathrm{c2}}}}\left[\mathrm{ln}{\displaystyle \frac{j_c}{j}}\right]^{2\sqrt{c_\mathrm{g}}},`$ (4.57) i.e., the system is a true superconductor also at $`T_\mathrm{g}`$. The transition to the high-temperature phase is therefore continuous at $`T_\mathrm{g}`$ where $`\rho (0)=0`$. The results of the RG calculation were derived under the assumption $`|\tau _\mathrm{g}|1`$ and no definite conclusions can be drawn for $`TT_\mathrm{g}`$. However, it is very likely that the dynamical critical exponent $`z`$ diverges as $`T`$ goes to zero. The increase of $`z`$ with decreasing $`T`$ was confirmed numerically by \[Lancaster and Ruiz-Lorenzo 1995\] (?). It is interesting to remark that one can obtain the current-voltage relation $`\dot{\phi }Ej^{z/2}`$ also from an argument which is a modification of a scaling argument first used by \[Fisher et al. 1991a\] (?). \[Fisher et al. 1991a\] (?) assume that the vortex-glass transition is continuous and characterized by the divergence of the characteristic length and time scales, $`\xi _\mathrm{g}`$ and $`t_{\xi _\mathrm{g}}\xi _\mathrm{g}^z`$, respectively. Their original argument is as follows: We express $`E`$ as $`\dot{𝐀}/c`$. Since in the GL-Hamiltonian $`𝐀`$ appears in the combination $`(\mathbf{}(e/c)𝐀)`$, $`𝐀`$ should then scale at the glass transition as an inverse length $`\xi _\mathrm{g}^1`$ and hence $`𝐄`$ as $`\dot{𝐀}\xi _\mathrm{g}^1t_{\xi _\mathrm{g}}^1\xi _\mathrm{g}^{1z}`$. If the critical properties are described by a finite-temperature fixed point, the free energy density $`𝐣𝐀`$ scales as $`\xi _\mathrm{g}^d`$ and hence $`j\xi _\mathrm{g}^{d+1}`$. The current-voltage relation has, therefore, in the critical region the form $`E\xi _\mathrm{g}^{1+z}=\mathrm{\Phi }_d(j\xi _\mathrm{g}^{d1}).`$ (4.58) Since $`\xi _\mathrm{g}`$ has to drop out of this relation for $`TT_\mathrm{g}`$, we obtain $`\mathrm{\Phi }_d(u)u^{(1+z)/(d1)}`$ and hence $`Ej^{(1+z)/(d1)}`$ (4.59) at $`T_\mathrm{g}`$, which is the result of \[Fisher et al. 1991a\] (?). In the present case of a two-dimensional system with $`𝐁`$ parallel to the plane this argument has to be modified. Since only the $`z`$ component $`B_z=_xA_y_yA_x`$ of the magnetic field is non-zero and because of the absence of any $`y`$ dependence in this geometry, the relevant component of $`𝐀`$ is $`A_y`$. In particular, there is no $`(_y\frac{e}{c}A_y)`$ term in $`_{\mathrm{GL}}`$ and hence we conclude $`A_y\xi _\mathrm{g}^0`$, $`E\xi _\mathrm{g}^z`$, $`j\xi _\mathrm{g}^2`$ which results in $`E\xi _\mathrm{g}^z=\mathrm{\Phi }_2(j\xi _\mathrm{g}^2).`$ (4.60) At the transition, this gives $`Ej^{z/2}`$, in agreement with the RG result $`Ej`$ (apart from a logarithmic correction) at the vortex glass transition temperature, where $`z=2`$. In the high-temperature phase ($`\tau _\mathrm{g}<0`$) where $`\mathrm{\Phi }_2j\xi _\mathrm{g}^2`$ and $`z=2`$, we get $`Ej`$, such that the correlation length drops out of the result. The same is true for $`T<T_\mathrm{g}`$, where the RG result gives $`Ej^{z/2}`$. From this one has to conclude that both phases are critical such that the transition is not accompanied by a diverging correlation length. For completeness, we remark here that \[Toner 1991a\] (?) and \[Nattermann et al. 1991\] (?) tried to calculate the current-voltage relation from an estimate of the energy barrier $`E_B(j)`$ between different meta-stable states. Both authors use the result from the statics for $`\sigma (L)`$, equation (4.29). While \[Toner 1991a\] (?) finds $`E_B(j)\tau _\mathrm{g}[\mathrm{ln}(j_c/j)]^{1/2}`$, \[Nattermann et al. 1991\] (?) get $`E_B(j)\tau _\mathrm{g}^2\mathrm{ln}(j_c/j)`$. Together with the general creep formula (3.73), the latter result gives a power law for the current-voltage relation, but with a wrong result for the exponent $`z2`$, which is proportional to $`\tau _\mathrm{g}`$ and not to $`\tau _\mathrm{g}^2`$. This discrepancy is related to the fact that for $`TT_\mathrm{g}`$ the physics is described by a finite-temperature fixed point at which energy barriers are not well defined. We conclude from this section that below the temperature $`T_\mathrm{g}`$ the vortex-line array in an impure superconducting film in a parallel field exhibits a glassy phase with vanishing linear resistivity. This phase is the most accurately studied example of a vortex glass (although of mainly academic interest). The spatial correlations of the vortex-line array exhibit a decay slightly faster than algebraic. The vortex-glass transition is not accompanied by a diverging correlation length (as assumed in the scaling argument of \[Fisher et al. 1991a\] (?) and as seen in many experiments for bulk materials) since both phases are critical. Because of the absence of shear modes, below $`T_\mathrm{g}`$ the system is both a positional and a phase-coherent vortex glass (in the sense discussed in section 2.4). ## 5 Superconducting film in a perpendicular field So far we examined in chapter 3 a single vortex as an example of a directed manifold and in chapter 4 a periodic array of vortex lines in a film. Both cases are simple in the sense that the vortex conformations can be described by a single-valued displacement field, or in other terms, that these systems are free of topological defects. The possible presence of such defects implies that the elastic approximation, which was the basis of our analysis in the previous chapters, may break down. We now turn to the simplest system that allows for topological defects: an array of point-vortices in a superconducting film induced by a perpendicular magnetic field. We consider here only the case of a thin film, where the film thickness $`s`$ is much smaller than the bulk correlation length $`\xi `$. Then the condensate wave function can be considered as constant in the direction perpendicular to the film, i.e., it is a two-dimensional (2D) degree of freedom. Nevertheless, the screening mechanism remains three-dimensional, since the electromagnetic field still propagates into the third dimension. ### 5.1 Pure system To set the stage for an analysis of pinning effects in this two-dimensional system, we briefly summarize important properties of the pure system (see in particular \[Doniach and Hubermann 1979\] (?), \[Hubermann and Doniach 1979\] (?), \[Fisher 1980\] (?)). The structure and interaction of vortices in thin films were first studied by Pearl (?, ?). Several distinct features are to be mentioned. (i) Because of screening, the creation of a vortex costs a finite amount of energy. This is in contrast to vortices in 2D superfluids, the energy of which diverges logarithmically with the system size due to the absence of screening. (ii) Every vortex has a magnetic moment $`M`$ that diverges proportional to the system size $`L`$. Thus the lower critical field vanishes in infinite films, $`H_{\mathrm{c1}}1/L`$. (iii) The vortex pair interaction potential is $`U(𝐱)`$ $`=`$ $`2s\epsilon _0U^{(0)}(x/L_\mathrm{s})`$ (5.1a) $`U^{(0)}(x)`$ $``$ $`\{\begin{array}{cc}\mathrm{ln}x,\hfill & x1\hfill \\ \frac{1}{x},\hfill & x1\hfill \end{array}`$ (5.1d) where $`s\epsilon _0s\left({\displaystyle \frac{\mathrm{\Phi }_0}{4\pi \lambda }}\right)^2`$ (5.2) is the energy scale and $`\lambda `$ is the bulk penetration length. A further important length scale is the screening length $`L_\mathrm{s}2\lambda ^2/s,`$ (5.3) which may be macroscopically large (in the range of millimeters). Screening sets in only on scales beyond $`L_\mathrm{s}`$, i.e., this scale is intimately related to the inhomogeneity of $`B`$. The importance of the screening length becomes evident already at zero magnetic field ($`H=0`$). Only in the limiting case $`L_\mathrm{s}\mathrm{}`$ the vortex self energy diverges logarithmically with the system size and individual vortices cannot be created by thermal activation. Then vortex-antivortex pairs interact logarithmically on large scales and dissociate according to the entropy-driven Kosterlitz-Thouless (KT) mechanism \[Kosterlitz and Thouless 1973\] at a temperature $`T_{\mathrm{KT}}{\displaystyle \frac{s\epsilon _0}{2}}={\displaystyle \frac{\mathrm{\Phi }_0^2}{16\pi ^2L_\mathrm{s}}}.`$ (5.4) In this expression we neglect renormalization effects which lead only to a quantitative reduction of the transition temperature. However, in the limit $`L_\mathrm{s}\mathrm{}`$ the transition temperature vanishes like $`T_{\mathrm{KT}}1/L_\mathrm{s}`$ and the system becomes critical only at zero temperature. At large but finite $`L_\mathrm{s}`$ a crossover in the vicinity of the small but finite $`T_{\mathrm{KT}}`$ survives. The electric resistivity of the film can be understood in terms of vortex-antivortex pairs under the influence of a transport current which drives vortices and antivortices in opposite directions \[Halperin and Nelson 1979\]. For finite $`L_\mathrm{s}`$ and $`T>0`$ the film has Ohmic resistivity at sufficiently small current densities, since vortex-antivortex pairs have to overcome only a finite energy barrier to dissociate. Thermal activation therefore leads to a finite dissociation rate and a resistivity proportional to the driving current. At larger current densities, where the vortex interaction is probed on scales $`rL_\mathrm{s}`$ (i.e., on this scale the attractive vortex interaction is balanced by the driving force), the logarithmic vortex interaction implies a power law for the current-voltage relation, which was predicted \[Halperin and Nelson 1979\] and observed experimentally \[Epstein et al. 1981, Kadin et al. 1983\] already long ago. More recent experimental \[Repaci et al. 1996\], numerical \[Simkin and Kosterlitz 1997\], and analytical \[Pierson and Valls 1999\] work was devoted to examining finite-size effects on the KT transition, which includes the case of a superconducting film regarding $`L_\mathrm{s}`$ as ‘intrinsic’ finite-size scale. Although the case of zero field would be of interest on its own, we subsequently consider only finite fields, $`H>0`$, which induce a finite average vortex density. Then one can think of the vortex system as a superposition of a neutral subsystem composed of vortex dipoles and a subsystem of excess vortices oriented in the direction parallel to $`𝐇`$. We now focus on the subsystem of excess vortices and assume that the effect of the dipole subsystem amounts to a screening of the interaction between the excess vortices. This interaction is expected to be purely repulsive even in the presence of screening by dipoles such that the ground state of the excess vortices should be a triangular lattice. Thermal fluctuations lead to the creation and annihilation of extra excess vortices and thereby create vacancies and interstitials in the lattice. Such fluctuations cost finite minimum energy in contrast to vortex displacements. Therefore we restrict the subsequent analysis to displacement fluctuations and examine their influence on the positional and orientational order of the lattice. In the literature the vortex lattice is, in general, not examined on the basis of the physically given vortex interaction (LABEL:V.Pearl). Instead, simplified model interactions are used. Since certain physical features depend on the choice of the model interaction, we attempt to classify them by distinguishing the following cases according to the behaviour of the potential on large length scales: $`U`$ is short-ranged, such that $`d^2xU(𝐱)<\mathrm{}`$. Only in this case the interaction energy of the vortex lattice is extensive and all elastic constants are finite. The interaction $`U`$ decays on large length scales but it is long-ranged, such that $`d^2xU(𝐱)=\mathrm{}`$ and the interaction energy is not extensive. This case includes the Pearl interaction. The system is incompressible, i.e., it has a singular compression modulus. The interaction is logarithmic on large length scales, $`U(𝐱)\mathrm{ln}x`$. This case, which also represents an incompressible system, includes the so-called two-dimensional Coulomb gas and the superconducting film in the limiting case when $`L_\mathrm{s}`$ is larger than the system size. We now briefly summarize the elastic properties of the system and the presence of phase transitions in these cases. Short-range case. Among these three cases, this one has been examined most since it can be represented by a well-defined elasticity theory with finite elastic constants. In general, the elastic Hamiltonian can be written in Fourier space as $`_{\mathrm{el}}[𝐮]={\displaystyle \frac{1}{2}}{\displaystyle _{BZ}}{\displaystyle \frac{d^2q}{(2\pi )^2}}[c_{11}(𝐪)q^2|𝐮_L(𝐪)|^2+c_{66}(𝐪)q^2|𝐮_T(𝐪)|^2].`$ (5.5) The displacement field is defined in position space only as a function of the discrete positions $`𝐗`$ of the undistorted lattice. In Fourier space it reads $`𝐮(𝐪)`$ $`=`$ $`a^2{\displaystyle \underset{𝐗}{}}e^{i𝐪𝐗}𝐮(𝐗)`$ (5.6) with the area $`a^2=\mathrm{\Phi }_0/B`$ per vortex. The displacement can be decomposed into a longitudinal and a transverse component $`𝐮_L(𝐪)=𝐏^L(𝐪)𝐮(𝐪),𝐮_T(𝐪)=𝐏^T(𝐪)𝐮(𝐪)`$ (5.7) by means of the projectors $`P_{\alpha \beta }^L(𝐪)=q_\alpha q_\beta /q^2`$ and $`P_{\alpha \beta }^T(𝐪)=\delta _{\alpha \beta }P_{\alpha \beta }^L(𝐪)`$. The energy of longitudinal and transverse displacements is given by the compression modulus $`c_{11}`$ and the shear modulus $`c_{66}`$, respectively. In general both moduli depend on the wave vector $`𝐪`$ because of the non-locality of the interaction. The elastic moduli are known explicitly for the special interaction $`U^{(0)}(x)=\mathrm{K}_0(x)`$ with the Bessel function $`\mathrm{K}_0`$, which decays exponentially beyond the screening length. This is the interaction of straight vortex lines in a three-dimensional superconductor apart from a replacement of the screening length of the bulk material, $`\lambda `$, by the screening length of the film, $`L_\mathrm{s}`$. Thus one finds from the bulk moduli \[Blatter et al. 1994\] within the continuum isotropic approximation $`c_{66}(𝐪)`$ $`=`$ $`2s\epsilon _0{\displaystyle \frac{B}{8\mathrm{\Phi }_0}},`$ (5.8a) $`c_{11}(𝐪)`$ $`=`$ $`2s\epsilon _0{\displaystyle \frac{2\pi L_\mathrm{s}^2}{1+L_\mathrm{s}^2q^2}}{\displaystyle \frac{B^2}{\mathrm{\Phi }_0^2}}.`$ (5.8b) In this chapter we use $`B`$ to parameterize the vortex density $`B/\varphi _0`$ per unit area rather than as average magnetic induction. In the case with finite elastic constants, the pure system can have two distinct phase transitions (for a review, see e.g. \[Strandburg 1987\] (?)): a melting transition between a solid and a hexatic liquid phase, and a transition between the hexatic and an isotropic liquid \[Kosterlitz and Thouless 1973, Halperin and Nelson 1978, Nelson 1978, Young 1979, Nelson and Halperin 1979\]. In this scenario, called the KTHNY scenario after the aforementioned authors, the solid has quasi-long-range positional order and long-range orientational order; the hexatic liquid has short-range positional order but still quasi-long-range orientational order, and finally in the isotropic liquid orientational order is also short-ranged. These transitions have been discussed in the particular context of superconducting films by \[Doniach and Hubermann 1979\] (?), \[Hubermann and Doniach 1979\] (?) and \[Fisher 1980\] (?). The melting transition is driven by the entropic unbinding of dislocation pairs. In a triangular lattice, which is the type of lattice formed by vortices, dislocations have a pair interaction $`U_{\mathrm{disloc}}(x){\displaystyle \frac{2}{\pi \sqrt{3}}}{\displaystyle \frac{c_{66}(c_{11}c_{66})}{c_{11}}}\mathrm{ln}(x/a)`$ (5.9) \[Young 1979, Fisher 1980\], where $`c_{ii}c_{ii}(𝐪=\mathrm{𝟎})`$. The given expression represents only the leading term for large distances; at smaller distances additional terms are present. Thus, according to the Kosterlitz-Thouless scenario, dislocation pairs unbind at a temperature $`T_\mathrm{m}{\displaystyle \frac{1}{2\pi \sqrt{3}}}{\displaystyle \frac{\mathrm{\Phi }_0}{B}}{\displaystyle \frac{c_{66}(c_{11}c_{66})}{c_{11}}},`$ (5.10) where we have ignored the renormalization of the elastic constants. The quantitatively correct value of $`T_\mathrm{m}`$ is hard to determine analytically, since it depends on the dislocation core energy, i.e. the discrete vortex structure of the core, which is difficult to include in analytic treatments. It is interesting to note that the melting transition can be captured by a modified Lindemann criterion. The traditional Lindemann criterion $`𝐮^2(𝐗)c_{\mathrm{Li}}^2a_{\mathrm{}}^2`$ for the stability of the lattice is not applicable, since in two dimensions $`𝐮^2(𝐗)=\mathrm{}`$ at any finite temperature. However, considering the relative displacement of neighbouring vortices (separated by a basis vector $`𝐛`$ with $`|𝐛|=a_{\mathrm{}}`$) instead of the absolute displacement, one obtains $`c_{\mathrm{Li}}^2a_{\mathrm{}}^2`$ $``$ $`[𝐮(𝐗)𝐮(𝐗+𝐛)]^2`$ (5.11) $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle _𝐪}(𝐪𝐛)^2W_{\alpha \alpha }(𝐛)`$ $``$ $`{\displaystyle \frac{a_{\mathrm{}}^2T}{4a^2c_{66}}},`$ where we have assumed $`c_{11}c_{66}`$ \[Scheidl and Vinokur 1998a\]. A comparison of (5.10) and (5.11) allows the determination of the Lindemann parameter $`c_{\mathrm{Li}}\left({\displaystyle \frac{1}{8\pi \sqrt{3}}}\right)^{1/2}0.15.`$ (5.12) This value may be a good reference value to estimate the stability of higher-dimensional systems with quasi-long-range positional order (such as the bulk vortex lattice with pinning) to the proliferation of topological defects. The second transition from a hexatic liquid to an isotropic liquid is driven by the unbinding of disclination pairs. Such pairs also have a logarithmic interaction, $`U_{\mathrm{discl}}(x)={\displaystyle \frac{\pi K_A}{18}}\mathrm{ln}(x/a)`$ (5.13) with the Frank constant $`K_A`$ \[Nelson and Halperin 1979\]. Thus disclination pairs unbind according to the KT mechanism at a temperature $`T_\mathrm{h}{\displaystyle \frac{\pi K_A}{72}}.`$ (5.14) Again, the quantitative value of $`T_\mathrm{h}`$ is very difficult to determine analytically. Partially, the difficulty is due to the strong temperature dependence of $`K_A=K_A(T)`$, which diverges above the melting transition. However, this divergence ensures that $`T_\mathrm{h}T_\mathrm{m}`$, i.e., that the lattice is more stable against disclinations than against dislocations. Thus, in order to study the stability of the elastic (solid) phase, it is sufficient to focus on the proliferation of dislocations. Long-range case. The actual interaction (LABEL:V.Pearl) between vortices in a film is long-ranged. If we ignore the regime on length scales below $`L_\mathrm{s}`$ and take $`U(x)=2s\epsilon _0L_\mathrm{s}/x`$, the interaction is the electrostatic one for a Wigner crystal. In this case the elastic moduli are \[Bonsall and Maradudin 1977\] $`c_{66}(𝐪)`$ $``$ $`0.252s\epsilon _0L_\mathrm{s}\left({\displaystyle \frac{B}{\mathrm{\Phi }_0}}\right)^{3/2},`$ (5.15a) $`c_{11}(𝐪)`$ $`=`$ $`2s\epsilon _0L_\mathrm{s}{\displaystyle \frac{2\pi }{q}}\left({\displaystyle \frac{B}{\mathrm{\Phi }_0}}\right)^2.`$ (5.15b) As in the short-ranged case, the shear modulus is essentially local and, in particular, $`c_{66}(𝐪=0)`$ is finite. However, $`c_{11}(𝐪)q^1`$, which means that the vortex lattice is incompressible. Thus conventional local elasticity theory does not apply. Nevertheless, by formally taking $`c_{11}\mathrm{}`$, most of the results obtained for the short-ranged case can be transferred to this long-ranged case. Since the dislocation interaction remains logarithmic on large scales one still expects a melting transition at a finite temperature according to equation (5.10). Logarithmic case. Finally, the logarithmic case is of interest, since it describes the superconducting film in the limit $`L_\mathrm{s}\mathrm{}`$. In this limit, the elastic moduli can be obtained directly from equations (LABEL:2d.moduli), $`c_{66}(𝐪)`$ $`=`$ $`2s\epsilon _0{\displaystyle \frac{B}{8\mathrm{\Phi }_0}},`$ (5.16a) $`c_{11}(𝐪)`$ $`=`$ $`2s\epsilon _0{\displaystyle \frac{2\pi }{q^2}}{\displaystyle \frac{B^2}{\mathrm{\Phi }_0^2}}.`$ (5.16b) Again, $`c_{66}`$ is finite on large scales, but the longer range of the interaction has even further increased the incompressibility. Nevertheless, a melting transition should be present according to the formal argument used already for the long-ranged case. The analytical description of the two distinct transitions (solid-to-liquid and hexatic-to-isotropic) assumes that the topological defects are very dilute at the transition, i.e., that they have a sufficiently large core energy. However, for small core energies the KT transition – which is thermodynamically of infinite order – may become a first-order transition (FOT) \[Minnhagen 1987, Thijssen and Knops 1988\]. Thus it is not clear whether the 2D melting must have a unique nature or whether there can to be two separate transitions for the loss of positional and orientational order. There has been controversial evidence for both the KTHNY scenario and a FOT scenario. Whereas earlier simulations always seemed to support the FOT scenario, it is now believed that this might be an artifact for finite-size effects and more recent simulations \[Bagchi et al. 1966, Pérez-Garrido and Moore 1998, Alonso and Fernandez 1999, Jaster 1999\] provide evidence for the KTHNY scenario, which seems to be particularly well established for particles with short-ranged interactions (see, e.g., \[Chaikin and Lubensky 1995\] (?)). For the long-ranged case and even more for the logarithmic case, the thermodynamic nature of the transition is not unambiguously understood. Conventional thermodynamics already runs into trouble since the energy of the system is no longer extensive. To close this section on the pure system, we quote some references to provide the reader with some references into the literature concerning the long-ranged systems rather than to give a comprehensive overview. \[Calliol et al. 1982\] (?) examine the long-ranged case with $`U(x)1/x`$ and find crystallization at $`T\frac{1}{140}2s\epsilon _0`$ (without being able to resolve the question of whether it is a single transition or a sequence of two transitions). This value of $`T_\mathrm{m}`$ is smaller than the one given in equation (5.10) by a factor of approximately 5, which in principle can be ascribed to a renormalization of elastic constants that is not taken into account in equation (5.10). Recent simulations \[Moore and Pérez-Garrido 1999\] of the logarithmic case suggest that the solid might be destroyed at any finite $`T`$. Many simulations have been performed on the model in the lowest Landau-level (LLL) approximation, which also results in a logarithmic vortex interaction due to the absence of screening. On the one hand there is evidence suggestive of a first-order melting \[Hu and MacDonald 1993, Kato and Nagaosa 1993b, Šašik and Stroud 1994b\] using the LLL approximation, but on the other hand there are also claims of the absence of the solid phase at finite temperatures (\[O’Neill and Moore 1992\] ?, ?, \[Yeo and Moore 1996a\] ?, ?, ?). \[Moore and Pérez-Garrido 1999\] (?) argue that in the logarithmic case dislocations screen the interaction of disclinations such that disclination pairs unbind and destroy crystalline order at any finite temperature. However, in the long-range case numerical simulations are likely to be affected by finite-size effects and the choice of boundary conditions even for very large system sizes. On the experimental side there is evidence for the KTHNY scenario in superconducting films \[Wördenweber et al. 1986, Berghuis et al. 1990, Theunissen et al. 1996\]. In other systems experimental evidence supports the FOT scenario (see \[Chaikin and Lubensky 1995\] (?) and \[Pérez-Garrido and Moore 1998\] (?) for a summary), which partially may be ascribed to the presence of additional symmetry-breaking fields. ### 5.2 Disordered vortex lattice without dislocations We now address the question again of how quenched impurities affect the solid vortex lattice. Hereby we focus on the case with short-ranged interactions. Then the vortex interactions can be represented by the elastic Hamiltonian (5.5), provided displacements are small (i.e., the phase is solid). Since we aim mainly at the properties of the system on length scales larger than $`L_\mathrm{s}`$, we may ignore the dispersion of the elastic constants, the value of which will be renormalized by fluctuations on small scales, as argued in section 4. Although the incompressible vortex lattice in a film does not strictly belong to the short-ranged case, it should be well captured by the formal limit $`c_{11}\mathrm{}`$. The effect of disorder on the solid vortex lattice of the film in a perpendicular field turns out to be very similar to that of the film in a parallel field, as treated in the previous chapter. The main complication consists in the fact that the displacement field has two (instead of one) components and, consequently, the different elastic response of shear and compression modes. Our presentation of this generalization closely follows \[Carpentier and Le Doussal 1997\] (?) and \[Carpentier 1999\] (?). According to the lines sketched in Appendix A the effective contribution of pinning to the Hamiltonian can be written in the form $`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle \underset{𝐗}{}}V(𝐗+𝐮(𝐗))`$ (5.17) $`=`$ $`{\displaystyle d^2x\left\{\stackrel{~}{V}(𝐮(𝐱),𝐱)\frac{1}{2}\mu _{\alpha \beta }(𝐱)[_\alpha u_\beta (𝐱)+_\beta u_\alpha (𝐱)]\right\}}`$ with an effective potential $`\stackrel{~}{V}(𝐮(𝐱),𝐱)=\rho _𝐮(𝐱)V(𝐱)`$ of zero average and $`\overline{\stackrel{~}{V}(𝐮,𝐱)\stackrel{~}{V}(𝐮^{},𝐱^{})}`$ $``$ $`\stackrel{~}{\mathrm{\Delta }}(𝐮𝐮^{})\delta (𝐱𝐱^{}),`$ (5.18a) $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ $`=`$ $`\rho _0{\displaystyle \underset{𝐗}{}}\mathrm{\Delta }(𝐗+𝐮),`$ (5.18b) where the effective correlator $`\stackrel{~}{\mathrm{\Delta }}`$ is periodic due to a sum over the undistorted lattice positions $`𝐗`$ and $`\mathrm{\Delta }(𝐱)=\overline{V(𝐱)V(\mathrm{𝟎})}`$ is the correlator of the original pinning potential. From the correlator of $`\stackrel{~}{V}`$ we have separated a random field $`𝝁`$, the correlation of which can be written as $`\overline{\mu _{\alpha \beta }(𝐱)\mu _{\gamma \delta }(𝐱^{})}`$ $`=`$ $`[(\sigma _{11}c_{11}^22\sigma _{66}c_{66}^2)\delta _{\alpha \beta }\delta _{\gamma \delta }`$ (5.19) $`+\sigma _{66}c_{66}^2(\delta _{\alpha \gamma }\delta _{\beta \delta }+\delta _{\alpha \delta }\delta _{\beta \gamma })]Q_{\mathrm{}}^2\delta (𝐱𝐱^{}).`$ For the unrenormalized model $`\sigma _{66}=0`$ and $`\sigma _{11}=Q_{\mathrm{}}^2\rho _0^2\mathrm{\Delta }_0/T^2`$ arises from the coupling of the random potential $`V`$ to the divergence of the displacement field (cf. equation (A.4)). The length $`Q_{\mathrm{}}`$ of the smallest non-vanishing reciprocal lattice vector (RLV) is related to the vortex spacing via $`Q_{\mathrm{}}^2=16\pi ^2/3a_{\mathrm{}}^2`$. Equation (5.17) is the generalized analog of equation (4.17) for the film in a parallel field. \[Carpentier and Le Doussal 1997\] (?) and \[Carpentier 1999\] (?) studied this model retaining only the contribution from the smallest non-vanishing RLV to the correlator, i.e., approximating $`\stackrel{~}{\mathrm{\Delta }}(𝐮)\rho _0^2_{𝐐:|𝐐|=Q_{\mathrm{}}}`$ $`\widehat{\mathrm{\Delta }}(𝐐)\mathrm{cos}[𝐐𝐮]`$. Then the Fourier coefficients $`\widehat{\mathrm{\Delta }}(𝐐)`$ act like a random-field amplitude, cf. equation (4.18c), for which we introduce an effective variance with bare value $`g=\rho _0\widehat{\mathrm{\Delta }}(Q_{\mathrm{}})/T^2`$. Under renormalization, not only the field amplitude $`g`$ evolves, but the random potential induces also bond disorder, i.e., a flow of $`\sigma _{11}`$ and $`\sigma _{66}`$. \[Carpentier and Le Doussal 1997\] (?) and \[Carpentier 1999\] (?) found the flow equations $`_lc_{11}`$ $`=`$ $`0,`$ (5.20a) $`_lc_{66}`$ $`=`$ $`0,`$ (5.20b) $`_l\sigma _{11}`$ $`=`$ $`b_{11}(\alpha )g^2,`$ (5.20c) $`_l\sigma _{66}`$ $`=`$ $`b_{66}(\alpha )g^2,`$ (5.20d) $`_lg`$ $`=`$ $`2\left(1{\displaystyle \frac{T}{T_\mathrm{g}}}\right)gb_\mathrm{g}(\alpha )g^2,`$ (5.20e) with the coefficients $`b_{11}(\alpha )`$ $``$ $`{\displaystyle \frac{3\pi }{4}}{\displaystyle \frac{T^2Q_{\mathrm{}}^4}{c_{11}^2}}[2I_0(\alpha )+I_1(\alpha )],`$ (5.21a) $`b_{66}(\alpha )`$ $``$ $`{\displaystyle \frac{3\pi }{4}}{\displaystyle \frac{T^2Q_{\mathrm{}}^4}{c_{66}^2}}[2I_0(\alpha )I_1(\alpha )],`$ (5.21b) $`b_\mathrm{g}(\alpha )`$ $``$ $`2\pi [2I_0(\alpha /2)I_0(\alpha )],`$ (5.21c) that depend through the Bessel functions $`I_{0,1}`$ on the parameter $`\alpha `$ $``$ $`{\displaystyle \frac{TQ_{\mathrm{}}^2}{4\pi }}\left({\displaystyle \frac{1}{c_{66}}}{\displaystyle \frac{1}{c_{11}}}\right),`$ (5.22) which measures the difference between the lattice response to shear and compression. From the linear term in equation (5.20e) one recognizes that disorder is relevant only at temperatures below the glass transition temperature $`T_\mathrm{g}`$. Its value is given by $`T_\mathrm{g}`$ $``$ $`{\displaystyle \frac{8\pi }{Q_{\mathrm{}}^2}}{\displaystyle \frac{c_{11}c_{66}}{c_{11}+c_{66}}}={\displaystyle \frac{\sqrt{3}}{\pi }}{\displaystyle \frac{\mathrm{\Phi }_0}{B}}{\displaystyle \frac{c_{11}c_{66}}{c_{11}+c_{66}}}`$ (5.23) \[Giamarchi and Le Doussal 1995\]. In principle, there is a finite renormalization of the elastic constants on small length scales because of the finiteness of the disorder correlation length $`\xi `$. On large scales, where the pinning energy correlator effectively factorizes as in equation (5.2), the renormalization of the elastic constants is absent due to a statistical symmetry analogous to the tile symmetry discussed in section 3.1.1d. The short-scale renormalization is neglected here, since it does not influence the large-scale properties of the vortex lattice. However, to be accurate the transition temperature is given by equation (5.23) using the renormalized value of the elastic constants. The fluctuations of the displacement field can be represented by the correlation function in the form $`W_{\alpha \beta }(𝐱)`$ $``$ $`\overline{[u_\alpha (𝐱)u_\alpha (\mathrm{𝟎})][u_\beta (𝐱)u_\beta (\mathrm{𝟎})]}`$ (5.24) $`=`$ $`W_L(𝐱)P_{\alpha \beta }^L(𝐱)+W_T(𝐱)P_{\alpha \beta }^T(𝐱)`$ with projectors $`P_{\alpha \beta }^L(𝐱)x_\alpha x_\beta /𝐱^2`$ and $`P_{\alpha \beta }^T\delta _{\alpha \beta }P_{\alpha \beta }^L`$. Above the glass transition, $`T>T_\mathrm{g}`$, the renormalization of the elastic constants and of the strengths $`\sigma _{11}`$ and $`\sigma _{66}`$ of the random ‘tilt’ fields are finite such that $`W(𝐱)\mathrm{ln}(x/a)`$ and the correlation function $`S(𝐐,𝐱𝐱^{})`$ decays algebraically, $`S(𝐐,𝐱𝐱^{})`$ $`=`$ $`\overline{e^{i𝐐[𝐮(𝐱)𝐮(𝐱^{})]}}\left({\displaystyle \frac{|𝐱𝐱^{}|}{a}}\right)^{\eta _Q},`$ (5.25a) $`\eta _Q`$ $`=`$ $`{\displaystyle \frac{Q^2}{2\pi Q_{\mathrm{}}^2}}\left({\displaystyle \frac{TQ_{\mathrm{}}^2}{c_{11}}}+{\displaystyle \frac{TQ_{\mathrm{}}^2}{c_{66}}}+\sigma _{11}+\sigma _{66}\right).`$ (5.25b) Thus there is quasi-long-range order as in the absence of disorder, and the effect of disorder (as seen in the static structure factor) amounts to an increased effective temperature. Below the glass transition, $`\tau _\mathrm{g}1T/T_\mathrm{g}>0`$, $`g`$ flows to a certain finite fixed-point value and generates an unlimited increase $`\sigma _{11}\sigma _{66}\mathrm{ln}(L/a)`$ with increasing length scale $`Le^l`$. Therefore the roughness is larger than logarithmic and, to leading order in $`\tau _\mathrm{g}`$, is given by \[Carpentier and Le Doussal 1997\]: $`W_T(𝐱)W_L(𝐱)`$ $``$ $`{\displaystyle \frac{w(\alpha )}{Q_{\mathrm{}}^2}}\tau _\mathrm{g}^2\mathrm{ln}^2(x/a),`$ (5.26a) $`W_T(𝐱)W_L(𝐱)`$ $``$ $`{\displaystyle \frac{\stackrel{~}{w}(\alpha )}{Q_{\mathrm{}}^2}}\tau _\mathrm{g}^2\mathrm{ln}(x/a),`$ (5.26b) $`w(\alpha )`$ $``$ $`6{\displaystyle \frac{2I_0(\alpha )(1+\alpha ^2/4)\alpha I_1(\alpha )}{[2I_0(\alpha /2)I_0(\alpha )]^2}},`$ (5.26c) $`\stackrel{~}{w}(\alpha )`$ $``$ $`6{\displaystyle \frac{I_1(\alpha )(1+\alpha ^2/4)2\alpha I_0(\alpha )}{[2I_0(\alpha /2)I_0(\alpha )]^2}}.`$ (5.26d) Precisely at the transition the parameter $`\alpha (T_\mathrm{g})`$ $``$ $`2{\displaystyle \frac{c_{11}c_{66}}{c_{11}+c_{66}}}`$ (5.27) depends on the ratio of the elastic constants only. The main structural feature, the increase $`W(𝐱)\mathrm{ln}^2(x/a)`$, is thus similar for magnetic fields parallel and perpendicular to the film, cf. equation (4.32). A further similarity is found in the dynamical properties, where a dynamical exponent \[Carpentier and Le Doussal 1997\] $`z=2+3e^{\gamma _{\mathrm{Eu}}}{\displaystyle \frac{(2+\alpha )[(2\alpha )/(2+\alpha )]^{(2\alpha )/4}}{2I_0(\alpha /2)I_0(\alpha )}}\tau _\mathrm{g}`$ (5.28) (with the Euler constant $`\gamma _{\mathrm{Eu}}`$) is found on the basis of the overdamped equation of motion $`\eta _0_tu_\alpha `$ $`=`$ $`{\displaystyle \frac{\delta }{\delta u_\alpha }}+\zeta _\alpha ,`$ (5.29a) $`\zeta _\alpha (𝐱,t)\zeta _\beta (𝐱^{},t^{})`$ $`=`$ $`2\eta _0T\delta _{\alpha \beta }\delta (𝐱𝐱^{})\delta (tt^{}).`$ (5.29b) Consequently, the vortex lattice responds with a velocity $`vF^{z/2}`$ (5.30) to a driving force $`F`$. Since $`F`$ is proportional to the current density and $`v`$ is proportional to the electric field, the linear electric resistivity vanishes in the glassy phase because of $`z>2`$ in agreement with the film in a parallel field, cf. equation (4.56). ### 5.3 Disordered vortex lattice with dislocations The analysis of the previous subsection was based on the elastic approximation, excluding topological defects such as dislocations. In order to check whether the glassy features found above actually describe the vortex system, it is crucial to examine the relevance of dislocations. By comparing the melting temperature and the glass-transition temperatures, equations (5.10) and (5.23), we immediately run into a dilemma because $`T_\mathrm{m}/T_\mathrm{g}[1(c_{66}/c_{11})^2]/6<1`$. At $`T>T_\mathrm{m}`$ the pure system is liquid and the effects of disorder cannot be studied starting from the elastic approximation (disorder will certainly not stabilize the ordered phase). On the other hand, if $`T<T_\mathrm{m}`$, then also $`T<T_\mathrm{g}`$ and disorder generates $`W(𝐱)\mathrm{ln}^2(x/a)`$. Unfortunately, this increased roughness due to the divergent renormalization of the bond disorder makes dislocation pairs unbind, because on large scales the gain of pinning energy is larger than the attraction energy of dislocation pairs. (The situation is analogous to the $`XY`$ model, where bond disorder beyond a finite critical strength induces vortices, see e.g. \[Nattermann et al. 1995\] (?).) Consequently, on length scales beyond a scale $`L_{\mathrm{disloc}}`$, where disorder induces the dissociation of dislocation pairs, dislocations are relevant at all temperatures and the elastic description breaks down and there can be neither a true melting transition nor a true glass transition. Only on length scales below $`L_{\mathrm{disloc}}`$ the system can be described within the elastic approximation. This scale is of the order of the vortex spacing slightly above $`T_\mathrm{m}`$ but it can become very large below $`T_\mathrm{m}`$ if disorder is very weak. In this case melting survives as a crossover phenomenon, i.e., as a rapid decrease of $`L_{\mathrm{disloc}}`$ if the temperature is increased above a renormalized $`T_\mathrm{m}^{}<T_\mathrm{m}`$. The length scale $`L_{\mathrm{disloc}}`$ can be estimated on the basis of the above RG flow equations \[Carpentier and Le Doussal 1998\]. For qualitative purposes, one can examine $`L_{\mathrm{disloc}}`$ from the simpler random-field $`XY`$ model. Using further approximations \[Giamarchi and Le Doussal 1995\] (?) and \[Le Doussal and Giamarchi 1998a\] (?) found that for temperatures slightly below $`T_\mathrm{m}`$ topological defects appear beyond $`L_{\mathrm{disloc}}R_ae^{c_1[(T_\mathrm{m}/T1)\mathrm{ln}(R_a/a)]^{1/2}}.`$ (5.31) Here $`R_a`$ is the length scale on which the displacement becomes of the order of the vortex spacing, $`W(R_a)=a^2`$, and $`c_1`$ is a numerical constant. The last equation overestimates $`L_{\mathrm{disloc}}`$ at low temperatures, where $`L_{\mathrm{disloc}}`$ saturates at a finite value for sufficiently weak disorder. This value is obtained from equation (5.31) after replacing $`T`$ by an effective value $`T^{}<T_\mathrm{m}`$ that depends only on the disorder strength \[Carpentier and Le Doussal 1998, Le Doussal and Giamarchi 1998a\]. Given the finiteness of $`L_{\mathrm{disloc}}`$ even at very low temperatures, one has to draw the conclusion that in a superconducting film with perpendicular field an elastic vortex glass, where dislocations would be irrelevant on asymptotically large scales, does not exist. Evidence for this fact stems also from numerical calculations \[Shi and Berlinski 1991, Gingras and Huse 1996, Middleton 1998, Zeng et al. 1999a\]. The resulting physical situation can be thought of as a vortex lattice broken into crystallites of size $`L_{\mathrm{disloc}}`$. As an immediate consequence of this fragmentation, the system has only short-ranged positional order (but the correlation length diverges with vanishing disorder strength for $`T<T_\mathrm{m}`$). A finite correlation length also implies finite pinning energies for these crystallites. Therefore thermal activation leads to a linear current-voltage relation for sufficiently weak currents. ## 6 Bulk Superconductor In this chapter we consider the weakly disordered bulk superconductor in an external field. We will first treat the system in the elastic approximation by neglecting dislocations. Later we show that this is indeed justified in a certain region of the phase diagram. ### 6.1 Elastic approximation For not too large fields, $`HH_{\mathrm{c2}}`$, we may use the London approximation to describe the vortex-line lattice (VLL). Then the elastic Hamiltonian can be written in Fourier space as $`_{\mathrm{el}}={\displaystyle \frac{1}{2}}{\displaystyle _{BZ}}{\displaystyle \frac{d^dq}{(2\pi )^d}}\{[c_{11}q_{}^2+c_{44}q_{}^2]|𝐮_L(𝐪)|^2+[c_{66}q_{}^2+c_{44}q_{}^2]|𝐮_T(𝐪)|^2\},`$ (6.1) involving only wave vectors $`𝐪=(𝐪_{},𝐪_{})`$ within the first Brillouin zone (BZ). The displacement field $`𝐮=𝐮_L+𝐮_T`$ has been decomposed into its longitudinal and transverse component with respect to $`𝐪_{}`$ as in equation (5.7). We extended the model from three to general dimension $`d`$ since we will perform below an expansion around $`d=4`$ dimensions below. In this generalization $`𝐪_{}`$ stands for the two-dimensional component of the wave vector perpendicular to the generalized magnetic field and $`𝐪_{}`$ stands for the $`(d2)`$-dimensional parallel component. In principle, the elastic constants are scale (momentum) dependent. However, since we are interested in the case of weak disorder, which becomes relevant only on large scales ($`L\lambda `$), we may neglect the dispersion of the elastic constants. In the case of strong disorder, in which relevant length scales (like the Larkin length) are smaller than $`\lambda `$, the consideration of the dispersion of the elastic constants can be achieved in principle (see \[Kierfeld 1998\] (?)). Next we include the interaction between the randomly distributed impurities and the vortex lattice. As in the previous chapters for simplicity we assume that the impurities result in a random potential $`V(𝐫)`$, which has zero average and is Gaussian correlated with two-point correlations $`\overline{V(𝐫)V(\mathrm{𝟎})}=\mathrm{\Delta }(𝐱)\delta ^{(d2)}(𝐳).`$ (6.2) The weight of this correlator is $`\mathrm{\Delta }_0=d^2x\mathrm{\Delta }(𝐱)=f_{\mathrm{pin}}^2n_\mathrm{i}^{(d)}\xi ^6`$, where $`f_{\mathrm{pin}}`$ denotes the pinning force of an individual impurity, $`n_\mathrm{i}^{(d)}`$ is the impurity density, and $`\xi `$ the maximum of the coherence and disorder correlation length \[Blatter et al. 1994\]. The pinning energy is then $`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle d^2xd^{d2}z\rho _𝐮(𝐫)V(𝐫)}`$ (6.3) $`=`$ $`{\displaystyle \underset{𝐗}{}}{\displaystyle d^{d2}zV(𝐗+𝐮(𝐗,𝐳),𝐳)}={\displaystyle d^dr\stackrel{~}{V}(𝐮(𝐫),𝐫)}`$ where $`\rho _𝐮(𝐫)=_𝐗\delta ^{(2)}(𝐱𝐗𝐮(𝐗,𝐳))`$ denotes the vortex-line density (per unit area) and $`𝐗`$ is a vector of the Abrikosov triangular lattice. $`\stackrel{~}{V}(𝐮(𝐫),𝐫)\rho _𝐮(𝐫)V(𝐫)`$ is the pinning energy density (per unit volume). It is interesting to note that the energy of the distorted VLL has certain symmetries. The pinning energy density $`\stackrel{~}{V}(𝐮,𝐫)\rho _𝐮(𝐫)V(𝐫)`$ is apparently invariant with respect to the transformation $`𝐮(𝐗,𝐳)𝐮(𝐗^{},𝐳)+𝐗^{}𝐗,`$ (6.4) i.e., by relabeling the vortex lines. Such relabeling also leaves the elastic Hamiltonian invariant and this symmetry is a symmetry of the total Hamiltonian $`=_{\mathrm{el}}+_{\mathrm{pin}}`$. $`_{\mathrm{el}}`$ possesses a second symmetry related to a rotation of the displacement field and a simultaneous exchange of shear and compression modulus. To demonstrate this symmetry, it is useful to introduce $`c_\mathrm{s}\sqrt{c_{11}c_{66}},\gamma {\displaystyle \frac{c_{66}}{c_{11}}}`$ (6.5) and to rescale $`𝐳𝐳^{}=𝐳\sqrt{c_\mathrm{s}/c_{44}}`$, which leads to $`_{\mathrm{el}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}c_\mathrm{s}\left({\displaystyle \frac{c_{44}}{c_\mathrm{s}}}\right)^{d2}{\displaystyle }d^2xd^{d2}z^{}\{\gamma ^{1/2}(\mathbf{}_{}𝐮)^2`$ (6.6) $`+\gamma ^{1/2}(\mathbf{}_{}𝐮)^2+(\mathbf{}_{}^{}𝐮)^2\}.`$ As far as the thermodynamic properties of the pure system are concerned, these are invariant under the transformation $`\gamma \gamma ^1`$, since the Hamiltonian is invariant under the simultaneous transformation ($`𝐮(u_1,u_2)`$) $`\gamma \gamma ^1,u_1u_1^{}=u_2,u_2u_2^{}=u_1`$ (6.7) (which also amounts to transforming the longitudinal displacement component into the transverse one and vice versa) and the partition function is invariant under the rotation of coordinates. The correlation functions will change sign if they include an odd power of $`u_2`$. We note, however, that the symmetry (6.7) breaks down as soon as dislocations are present, since the interaction energy of a dislocation pair is not symmetric with respect to the transformation $`\gamma \gamma ^1`$ and a full treatment including both $`c_{66}`$ and $`c_{11}`$ is required for studying the general case. In the presence of pinning this symmetry is absent even as a statistical symmetry in a strict sense. This can be seen from the disorder-averaged replicated pinning energy. However, we will find below that this symmetry is restored on large scales $`Ra`$. ### 6.2 The Bragg glass Historically, three different approximations have been discussed in treating the pinning Hamiltonian, which turn out to be valid on (i) small, (ii) intermediate and (iii) large length scales. (i) In his pioneering work on pinning in type-II superconductors, \[Larkin 1970\] (?) expanded (6.3) for small displacements $`𝐮`$ $`_{\mathrm{pin}}{\displaystyle d^{d2}z\underset{𝐗}{}\{V(𝐗,𝐳)+𝐮(𝐗,𝐳)\mathbf{}_{}V(𝐗,𝐳)+O(u^2)\}}.`$ (6.8) $`\mathbf{}_{}V(𝐗,𝐳)`$ is a random pinning force per length which acts on the vortex line of rest position $`𝐗`$. Although it was proved by \[Efetov and Larkin 1977\] (?) that, using perturbation theory, the lowest order (linear) term in $`𝐮`$ in the expansion of $`_{\mathrm{pin}}`$ (under certain conditions) already gives the exact result for the thermodynamic quantities, it became clear later that perturbation theory breaks down in random systems with many energy minima \[Villain and Séméria 1983, Fisher 1985a\]. Thus we can use the expansion (6.8) only as long as $`|𝐮|`$ is smaller than the typical distance between energy minima of $`V(𝐫)`$ which is of the order of $`\xi `$. This restricts the validity of the expansion (6.8) to length scales small compared to the Larkin length. In $`d=3`$ one finds from first-order perturbation theory \[Larkin 1970\] $`W(𝐫)`$ $`=`$ $`\overline{[𝐮(𝐱,z)𝐮(\mathrm{𝟎},0)]^2}`$ (6.9) $``$ $`{\displaystyle \frac{\rho _0\mathrm{\Delta }^{(2)}}{\pi c_{44}^{1/2}}}\left\{{\displaystyle \frac{1}{c_{11}^{3/2}}}(𝐱^2+z_l^2)^{1/2}+{\displaystyle \frac{1}{c_{66}^{3/2}}}(𝐱^2+z_t^2)^{1/2}\right\},`$ where we introduced $`z_t^2={\displaystyle \frac{c_{66}}{c_{44}}}z^2,z_l^2={\displaystyle \frac{c_{11}}{c_{44}}}z^2.`$ (6.10) The coefficient $`\rho _0\mathrm{\Delta }^{(2)}\frac{1}{2}\mathbf{}_{}^2\mathrm{\Delta }(𝐱)|_{𝐱=\mathrm{𝟎}}f_{\mathrm{pin}}^2n_\mathrm{i}^{(3)}\xi ^2/a^2`$ has the meaning of the density of the fluctuations of the pinning forces in the volume fraction occupied by the vortex lines. As already mentioned, this result is valid only as long as $`W(𝐱,z)\xi ^2`$, which defines the Larkin lengths $`L_\xi `$ and $`R_\xi `$ parallel and perpendicular to the magnetic field respectively: $`L_\xi `$ $``$ $`{\displaystyle \frac{\pi \xi ^2}{\rho _0\mathrm{\Delta }^{(2)}}}{\displaystyle \frac{c_{11}c_{44}c_{66}}{c_{11}+c_{66}}},`$ (6.11a) $`R_\xi `$ $``$ $`{\displaystyle \frac{\pi \xi ^2}{\rho _0\mathrm{\Delta }^{(2)}}}{\displaystyle \frac{c_{44}^{1/2}(c_{11}c_{66})^{3/2}}{c_{11}^{3/2}+c_{66}^{3/2}}}.`$ (6.11b) In general dimensions perturbation theory gives a roughness exponent \[cf. equation (3.15)\] $`\zeta =\zeta _{\mathrm{rf}}{\displaystyle \frac{4d}{2}}.`$ (6.12) From equation (6.9) one finds for the correlation function of translational order $`S(𝐐,𝐫)=\overline{e^{i𝐐[𝐮(𝐫)𝐮(\mathrm{𝟎})]}},`$ (6.13) where $`𝐐`$ is a reciprocal lattice vector, an exponential decay with the correlation length $`(a_{\mathrm{}}^2/4\pi ^2\xi ^2)L_\xi `$ and $`(a_{\mathrm{}}/4\pi \xi ^2)R_\xi `$, respectively. Note, however, that the regime of the true exponential decay is not reached because of the restrictions $`x<R_\xi `$ and $`z<L_\xi `$. (ii) On length scales larger than $`L_\xi `$ and $`R_\xi `$, respectively, the increase of $`W(𝐫)`$ will continue, but with a smaller roughness exponent since perturbation theory is known to overestimate the influence of disorder. As long as the displacement of the vortex lines is much smaller than the lattice spacing $`a`$ of the VLL (but larger than $`\xi `$), each vortex line sees its own random potential which cannot be reached by other vortex lines. Hence, since $`|𝐮𝐮^{}|a`$, $`\overline{V(𝐗+𝐮,𝐳)V(𝐗^{}+𝐮^{},𝐳^{})}`$ $`=`$ $`\mathrm{\Delta }(𝐗𝐗^{}+𝐮𝐮^{})\delta (𝐳𝐳^{})`$ (6.14) $``$ $`\mathrm{\Delta }_0\delta _{𝐗,𝐗^{}}\delta _\xi ^{(2)}(𝐮𝐮^{})\delta (𝐳𝐳^{}).`$ The pinning energy density $`\stackrel{~}{V}(𝐮,𝐫)\rho _0V(𝐱+𝐮,𝐳)`$ then obeys the approximate relation ($`\rho _0=B/\mathrm{\Phi }_0=a^2`$) $`\overline{\stackrel{~}{V}(𝐮,𝐫)\stackrel{~}{V}(𝐮^{},𝐫^{})}\rho _0\mathrm{\Delta }(𝐮𝐮^{})\delta _a^{(2)}(𝐱𝐱^{})\delta (𝐳𝐳^{}),`$ (6.15) which agrees with that of the random-manifold model of chapter 3 \[Feigel’man et al. 1989, Bouchaud et al. 1991, Bouchaud et al. 1992\]. In the random-manifold regime the roughness exponent $`\zeta _{\mathrm{rm}}`$ cannot be calculated exactly. A crude estimate is given by the Flory result $`\zeta _\mathrm{F}=(4d)/60.167`$ according to equation (3.20). \[Note that explicit values for exponents, which no longer depend on $`d`$, are given for $`d=3`$.\] \[Emig et al. 1999\] (?) calculated $`\zeta _{\mathrm{rm}}`$ from a functional renormalization-group treatment and found $`\zeta _{\mathrm{rm}}0.175`$ which varies only weakly with $`\gamma =c_{66}/c_{11}`$ (see below). The decay of $`S(𝐐,𝐫)\mathrm{exp}[Q^2W(𝐫)/2]`$ in this regime is therefore of stretched exponential form. (iii) Finally, for $`LL_aL_\xi (a/\xi )^{1/\zeta _{\mathrm{rm}}}`$ (or $`RR_aR_\xi (a/\xi )^{1/\zeta _{\mathrm{rm}}}`$) the vortex line displacement becomes of order $`a`$. Clearly, in the approximation for the perturbative regime and the manifold regime used so far, $`\stackrel{~}{V}(𝐮,𝐫)`$ does not fulfill the invariance property (6.4). However, it is precisely this property which determines the physics in the regime of very large length scales $`LL_a`$ ($`RR_a`$). In this regime the effect of disorder is very weak since displacements of vortex lines larger than $`a`$ are not very favourable because there is already a vortex line within a distance $`a`$ of any impurity. To keep this feature in our pinning Hamiltonian, we rewrite the vortex-line density as $`\rho _𝐮(𝐫)`$ $`=`$ $`{\displaystyle \underset{𝐗}{}}\delta (𝐱𝐗𝐮(𝐗,𝐳))={\displaystyle d^2x^{}\underset{𝐗}{}\delta (𝐗𝐱^{})\delta (𝐱𝐱^{}𝐮(𝐱^{},𝐳))}`$ (6.16) $`=`$ $`{\displaystyle d^2x^{}\delta (𝐱𝐱^{}𝐮(𝐱^{},𝐳))\rho _0\underset{𝐐}{}e^{i𝐐𝐱^{}}}`$ $``$ $`\rho _0{\displaystyle \frac{1}{|1+_\alpha u_\alpha |}}{\displaystyle \underset{𝐐}{}}e^{i𝐐[𝐱𝐮(𝐫)]}.`$ The pair correlator of the pinning energy density $`\stackrel{~}{V}(𝐮,𝐫)=\rho _𝐮(𝐫)V(𝐫)`$, which is the only quantity which enters the following calculation, is then $`\overline{\stackrel{~}{V}(𝐮,𝐫)\stackrel{~}{V}(𝐮^{},𝐫^{})}`$ $``$ $`\rho _0^2\mathrm{\Delta }(𝐱𝐱^{})\delta (𝐳𝐳^{}){\displaystyle \frac{1}{|1+_\alpha u_\alpha |}}{\displaystyle \frac{1}{|1+_\beta ^{}u_\beta ^{}|}}`$ (6.17) $`\times {\displaystyle \underset{𝐐,𝐐^{}}{}}e^{i𝐐(𝐱𝐱^{})+i(𝐐+𝐐^{})𝐱^{}i𝐐𝐮i𝐐^{}𝐮^{}}.`$ To exploit (6.17) further, a few remarks are in order: (i) The terms in the sum over $`𝐐^{}`$ with $`𝐐+𝐐^{}\mathrm{𝟎}`$ on the right hand side of (6.17) are rapidly oscillating on scales $`|\mathrm{\Delta }𝐱^{}|a`$ and therefore average to zero. (For weak disorder, where $`L_\xi a`$, this is even true in the perturbative regime.) (ii) The denominators lead to terms of the form $`\sigma (_\alpha u_\alpha ^a)(_\beta u_\beta ^b)`$ in the replica Hamiltonian (density). In $`d>2`$ dimensions $`\frac{d\sigma }{dl}=(2d)\sigma +\mathrm{}`$ such that these terms renormalize to zero for weak disorder. We therefore omit them in the following. (iii) We approximate $`\overline{\stackrel{~}{V}(𝐮,𝐫)\stackrel{~}{V}(𝐮^{},𝐫^{})}\stackrel{~}{\mathrm{\Delta }}(𝐮𝐮^{})\delta (𝐫𝐫^{}).`$ (6.18) Hence we get $`\stackrel{~}{\mathrm{\Delta }}(𝐮)=\rho _0^2{\displaystyle \underset{𝐐}{}}\widehat{\mathrm{\Delta }}(𝐐)e^{i𝐐𝐮}=\rho _0{\displaystyle \underset{𝐗}{}}\mathrm{\Delta }(𝐮+𝐗),`$ (6.19) where $`\widehat{\mathrm{\Delta }}(𝐐)`$ is the Fourier transform of $`\mathrm{\Delta }(𝐱)`$, $`\widehat{\mathrm{\Delta }}(𝐐)={\displaystyle d^2x\mathrm{\Delta }(𝐱)e^{i𝐐𝐱}}.`$ (6.20) Apparently, the correlator $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ is invariant under the transformation $`𝐮𝐮+𝐗`$, where $`𝐗`$ is an arbitrary vector of the Abrikosov lattice. If $`\mathrm{\Delta }(𝐱)(\mathrm{\Delta }_0/2\pi \xi ^2)\mathrm{exp}(𝐱^2/2\xi ^2)`$ then $`\widehat{\mathrm{\Delta }}(𝐐)=\mathrm{\Delta }_0\mathrm{exp}(𝐐^2\xi ^2/2)`$ and $`\stackrel{~}{\mathrm{\Delta }}^{(2)}={\displaystyle \frac{1}{2}}\mathbf{}_{}^2\stackrel{~}{\mathrm{\Delta }}(𝐮)|_{𝐮=\mathrm{𝟎}}={\displaystyle \frac{1}{2}}\rho _0^2{\displaystyle \underset{𝐐}{}}\widehat{\mathrm{\Delta }}(𝐐)𝐐^2\rho _0\mathrm{\Delta }_0/2\pi \xi ^4.`$ (6.21) To obtain the disorder averaged configuration of the VLL on a particular length scale, one has to take into account the renormalization of $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ by fluctuations on shorter length scales. This can be done systematically by a functional renormalization group (FRG) \[Fisher 1986a\] for the replica Hamiltonian $`_n`$ resulting from equations (2.16), (6.3), and (6.18). After the disorder averaging we obtain $`_n`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{a}{}}{\displaystyle d^2xd^{d2}z\left\{c_{11}(\mathbf{}_{}𝐮^a)^2+c_{66}(\mathbf{}_{}𝐮^a)^2+c_{44}(\mathbf{}_{}𝐮^a)^2\right\}}`$ (6.22) $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{ab}{}}{\displaystyle d^dr\stackrel{~}{\mathrm{\Delta }}(𝐮^a𝐮^b)}.`$ Because of the statistical invariance of the replica Hamiltonian under a shift of $`𝐮`$ by an arbitrary vector field inducing a compression, shear and/or tilt of the VLL, there is no renormalization of the elastic moduli \[Hwa and Fisher 1994b\]. Therefore the temperature obeys the exact flow equation $`dT/dl=(d2)T`$ leading to a $`T=0`$ fixed point for $`d>2`$. Notice, however, that in the original model, (6.1) with (6.3), the statistical invariance is not fulfilled exactly on length scales smaller than $`a`$, leading to a small renormalization $`c_{ii}\stackrel{~}{c}_{ii}`$ of the elastic constants which will be considered from now on as effective parameters (similar effects have been considered in some detail in section 4.1). This renormalization is negligible for the calculation of the leading asymptotic behaviour on largest scales. In the above approximation, $`_n`$ including the pinning energy is invariant under the transformation (6.7). Therefore this transformation will be present as symmetry in all thermodynamic quantities (such as the the displacement correlation) calculated from (6.22). The FRG was applied to (6.22) in the case of a scalar displacement field $`u`$ first by \[Giamarchi and Le Doussal 1994\] (?, ?). Here we follow closely the derivation of \[Emig et al. 1999\] (?) who treat the general case of a vector field $`𝐮`$. In the present FRG only coordinates are rescaled, $`𝐫\mathrm{exp}(dl)𝐫`$, to keep the cutoff $`\mathrm{\Lambda }`$ fixed with $`\mathrm{\Lambda }dl`$ the infinitesimal width of the momentum shell. Because of the dispersion of the elastic constants on scales smaller than the penetration depth $`\lambda `$, we have to choose here $`\mathrm{\Lambda }2\pi /\lambda `$. Fluctuations on smaller scales can be ignored if the Larkin length $`L_\xi `$ is much larger than $`\lambda `$, i.e., for weak disorder. For larger disorder one has to take into account the dispersion of $`\stackrel{~}{c}_{11}`$ and $`\stackrel{~}{c}_{44}`$, which will result in a more complicated cross-over but which will not affect the asymptotic behaviour (see also section IV.D in \[Giamarchi and Le Doussal 1995\] (?)). The flow equation for $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ can then be derived along the lines discussed in detail in Refs. \[Fisher 1986b, Balents and Fisher 1993\]. In contrast to previous cases, \[Emig et al. 1999\] (?) took into account the existence of a longitudinal and a transverse part in the elastic propagator. With the replacement $`{\displaystyle \frac{C}{2a^2}}\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ $``$ $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ (6.23a) $`C`$ $``$ $`{\displaystyle _{|𝐪|=\mathrm{\Lambda }}}\left\{{\displaystyle \frac{1}{(\stackrel{~}{c}_{11}q_{}^2+\stackrel{~}{c}_{44}q_{}^2)^2}}+{\displaystyle \frac{1}{(\stackrel{~}{c}_{66}q_{}^2+\stackrel{~}{c}_{44}q_{}^2)^2}}\right\}`$ (6.23b) $`=`$ $`{\displaystyle \frac{1}{8\pi ^2}}{\displaystyle \frac{1+\gamma }{\stackrel{~}{c}_{44}\stackrel{~}{c}_{66}}}\mathrm{\Lambda }^ϵ`$ one obtains to lowest order in $`ϵ=4d`$ a flow equation for the renormalized and rescaled correlator on scale $`L=\lambda e^l`$ (as in most of the previous chapters we suppress here the RG-flow variable $`l`$, i.e., we express $`\stackrel{~}{\mathrm{\Delta }}(𝐮,L)\stackrel{~}{\mathrm{\Delta }}_l(𝐮)`$ as $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$) $`{\displaystyle \frac{d\stackrel{~}{\mathrm{\Delta }}(𝐮)}{dl}}`$ $`=`$ $`ϵ\stackrel{~}{\mathrm{\Delta }}(𝐮)+{\displaystyle \frac{a_{\mathrm{}}^2}{2}}\{\left(_1^2\stackrel{~}{\mathrm{\Delta }}\right)^2+\left(_2^2\stackrel{~}{\mathrm{\Delta }}\right)^2+2\left(_1_2\stackrel{~}{\mathrm{\Delta }}\right)^2`$ (6.24) $`+2\stackrel{~}{\mathrm{\Delta }}^{(2)}(_1^2+_2^2)\stackrel{~}{\mathrm{\Delta }}{\displaystyle \frac{\delta }{4}}[(_1^2\stackrel{~}{\mathrm{\Delta }}_2^2\stackrel{~}{\mathrm{\Delta }})^2+4\left(_1_2\stackrel{~}{\mathrm{\Delta }}\right)^2]\}`$ with the dimensionless parameter $`\stackrel{~}{\mathrm{\Delta }}^{(2)}_1_1\stackrel{~}{\mathrm{\Delta }}(𝐮)|_{𝐮=\mathrm{𝟎}}=_2_2\stackrel{~}{\mathrm{\Delta }}(𝐮)|_{𝐮=\mathrm{𝟎}}`$ (6.25) and $`_1=/u_1`$ etc.. Both the last equality as well as $`_1_2\stackrel{~}{\mathrm{\Delta }}(𝐮)|_{𝐮=\mathrm{𝟎}}=0`$ follow from the requirement of hexagonal symmetry for $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$. The anisotropy parameter is given by $$\delta (\gamma )=1\frac{2\mathrm{ln}(\gamma )}{\gamma \gamma ^1}=1\frac{2}{C}_𝐪\frac{1}{(\stackrel{~}{c}_{11}q_{}^2+\stackrel{~}{c}_{44}q_{}^2)(\stackrel{~}{c}_{66}q_{}^2+\stackrel{~}{c}_{44}q_{}^2)},$$ (6.26) i.e., $`0\delta 1`$ for any ratio $`\gamma =\stackrel{~}{c}_{66}/\stackrel{~}{c}_{11}`$. In the special case $`\stackrel{~}{c}_{11}=\stackrel{~}{c}_{66}`$ (i.e., $`\delta =0`$) the flow equation (6.24) reduces to that of \[Balents and Fisher 1993\] (?). If $`\stackrel{~}{c}_{11}\mathrm{}`$, as often assumed for VLLs, $`\delta =1`$. To obtain the renormalized function $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ on large length scales $`L`$ including the fixed point $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}(𝐮)`$ for $`L\mathrm{}`$, one has to integrate equation (6.24). With the bare correlator of equation (6.19) showing the full symmetry of the triangular lattice – translation, sixfold rotational axis, six mirror lines – and the flow of equation (6.24) preserving these symmetries as it ought to, the set of possible solutions is restricted to functions with the full lattice symmetry on every length scale. One can solve (6.24) in a straightforward manner by rewriting the functional flow equation as a set of non-linear ordinary flow equations for the Fourier coefficients $`\widehat{\mathrm{\Delta }}(𝐐)`$, cf. equation (6.20). Rather than solving for the fixed point $`\widehat{\mathrm{\Delta }}_{\mathrm{}}(𝐐)`$ directly, \[Emig et al. 1999\] (?) numerically integrated the flow equations exploiting the remaining point group symmetries. It turns out that there is convergence to a fixed point (the pinning force correlator of which is illustrated in figure 8) from a large basin of attraction. Of special interest is the flow of $`\stackrel{~}{\mathrm{\Delta }}_l^{(2)}`$, since the renormalized propagator is proportional to $`{\displaystyle \frac{\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(q^1)}{q^4}}q^{d+2\zeta },`$ (6.27) where $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(q^1)=\stackrel{~}{\mathrm{\Delta }}_{l=\mathrm{ln}(\mathrm{\Lambda }/q)}^{(2)}(\mathrm{\Lambda }/q)^{d4},`$ (6.28) which determines the roughness exponent $`\zeta `$ of the VLL. As shown in figure 9, $`\stackrel{~}{\mathrm{\Delta }}_{l=\mathrm{ln}(\mathrm{\Lambda }/q)}^{(2)}`$ exhibits three scaling regions. (i) On scales $`L=q^1<L_\xi `$, $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{ln}(\mathrm{\Lambda }L)}^{(2)}(\mathrm{\Lambda }L)^{4d}`$ which reproduces the result of \[Larkin 1970\] (?) for the perturbative regime. (ii) In the region $`L_\xi <L<L_a`$, $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{ln}(\mathrm{\Lambda }L)}^{(2)}(\mathrm{\Lambda }L)^{2\zeta _{\mathrm{rm}}(\gamma )}`$, where $`\zeta _{\mathrm{rm}}(\gamma )`$ is the roughness exponent of the random-manifold regime. Numerically, $`\zeta _{\mathrm{rm}}(\gamma )`$ ranges from $`0.1737`$ for $`\gamma =0`$ to $`0.1763`$ for $`\gamma =1`$ and is continuously increasing in this interval (see figure 10). (iii) Finally, for $`L_a<L`$, $`\stackrel{~}{\mathrm{\Delta }}^{(2)}`$ approaches a fixed point $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}`$ which determines the asymptotic Bragg glass regime. The numerical value of $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}`$, which is of order $`ϵ=4d1`$, still depends on $`\gamma =c_{66}/c_{11}`$. With the numerical value for $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}`$ at hands, the explicit form of the displacement correlations $`W_{\alpha \beta }(𝐫)=\overline{[u_\alpha (𝐫)u_\alpha (\mathrm{𝟎})][u_\beta (𝐫)u_\beta (\mathrm{𝟎})]}`$ (6.29) in the Bragg-glass phase is given by \[Emig et al. 1999\] $`W_{11}(𝐫)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}(\gamma )a_{\mathrm{}}^2}{1+\gamma }}\{\mathrm{ln}\left({\displaystyle \frac{x^2+z_t^2}{L_a^2}}\right)+\gamma \mathrm{ln}\left({\displaystyle \frac{x^2+z_l^2}{L_a^2}}\right)`$ (6.30a) $`+{\displaystyle \frac{x_2^2x_1^2}{x^2}}[1\gamma {\displaystyle \frac{z_t^2}{x^2}}\mathrm{ln}(1+{\displaystyle \frac{x^2}{z_t^2}})+\gamma {\displaystyle \frac{z_l^2}{x^2}}\mathrm{ln}(1+{\displaystyle \frac{x^2}{z_l^2}})]\},`$ $`W_{12}(𝐫)`$ $`=`$ $`{\displaystyle \frac{2\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}(\gamma )a_{\mathrm{}}^2}{1+\gamma }}{\displaystyle \frac{x_1x_2}{x^2}}\{\gamma 1\gamma {\displaystyle \frac{z_l^2}{x^2}}\mathrm{ln}(1+{\displaystyle \frac{x^2}{z_l^2}})`$ (6.30b) $`+{\displaystyle \frac{z_t^2}{x^2}}\mathrm{ln}(1+{\displaystyle \frac{x^2}{z_t^2}})\},`$ and $`W_{22}(𝐫)`$ follows from $`W_{11}(𝐫)`$ by permuting $`\stackrel{~}{c}_{11}`$ and $`\stackrel{~}{c}_{66}`$. To lowest order in $`ϵ=4d`$, these correlations lead to the translational order correlation function $`S(𝐐,𝐫)`$ $``$ $`\overline{\mathrm{exp}(i𝐐[𝐮(𝐫)𝐮(\mathrm{𝟎})])}`$ (6.31) $``$ $`g_𝐐L_a^{\overline{\eta }_𝐐}(x^2+z_t^2)^{\frac{\overline{\eta }_𝐐}{2(1+\gamma )}}(x^2+z_l^2)^{\frac{\overline{\eta }_𝐐}{2(1+1/\gamma )}}`$ with the non-universal $`\gamma `$-dependent exponent $`\overline{\eta }_𝐐=\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}(\gamma )(a_{\mathrm{}}Q)^2`$ (6.32) and the geometrical prefactor $`g_𝐐`$ $`=`$ $`\mathrm{exp}\{{\displaystyle \frac{\stackrel{~}{\mathrm{\Delta }}_{\mathrm{}}^{(2)}(\gamma )(a_{\mathrm{}}Q)^2}{1+\gamma }}[(\widehat{𝐱}\widehat{𝐐})^2{\displaystyle \frac{1}{2}}]`$ (6.33) $`\times [(1{\displaystyle \frac{z_t^2}{x^2}}\mathrm{ln}(1+{\displaystyle \frac{x^2}{z_t^2}}))\gamma (1{\displaystyle \frac{z_l^2}{x^2}}\mathrm{ln}(1+{\displaystyle \frac{x^2}{z_l^2}}))]\},`$ which completely describes the angular dependencies of the translational order. Note that the factor $`g_𝐐`$ goes to $`1`$ in the limit $`z\mathrm{}`$. Therefore, in this limit the dependence of $`S(𝐐,𝐫)`$ on the reciprocal lattice vector $`𝐐`$ remains only in the exponent $`\overline{\eta }_𝐐`$. Moreover, it is interesting to note that the exponents of the algebraic decay in equation (6.31) depend on the elastic moduli as soon as $`z`$ is finite even without taking into account the non-universality of the exponent $`\overline{\eta }_𝐐`$ itself. If one ignores the non-trivial $`\gamma `$ dependence of $`\overline{\eta }_𝐐`$, in the case $`z=0`$ the above formulas reduce to those found in \[Giamarchi and Le Doussal 1994\] (?, ?). A logarithmic roughness of $`W(𝐫)`$ was already predicted earlier from scaling arguments \[Nattermann 1990\] and found also from a variational treatment with replica-symmetry breaking by \[Korshunov 1993\] (?) and \[Giamarchi and Le Doussal 1994\] (?, ?). However, this method is not able to capture the non-universality of $`\overline{\eta }_𝐐`$. The $`\gamma `$-dependence of the exponent $`\overline{\eta }_𝐐_{\mathrm{}}`$ for a smallest reciprocal lattice vector $`Q_{\mathrm{}}`$ is depicted in figure 10 and varies numerically from $`\overline{\eta }_𝐐_{\mathrm{}}=1.143`$ for $`\gamma =0`$ to $`\overline{\eta }_𝐐_{\mathrm{}}=1.159`$ for $`\gamma =1`$. In isotropic superconductors at low temperatures, where vortex lines interact via central forces, one has $`0\gamma 1/3`$. $`\gamma 1/3`$ for $`\lambda a`$, i.e., for fields close to $`H_{\mathrm{c1}}`$, and $`\gamma 0`$ for $`HH_{\mathrm{c2}}`$. For most of the field region $`\gamma \mathrm{\Phi }_0/16\pi \lambda ^2B`$. Thus, an increase of the external field from $`H_{\mathrm{c1}}`$ to $`H_{\mathrm{c2}}`$ should result in an increase of $`\overline{\eta }_𝐐`$ and a decrease of $`\zeta _{\mathrm{rm}}`$. At higher temperatures, where the vortex line interaction is renormalized considerably by thermal fluctuations, as well as in anisotropic superconductors, the above inequality for $`\gamma `$ may no longer be fulfilled. Clearly, in the latter case our starting Hamiltonian (6.22) would also have to be modified. The non-universality of $`\overline{\eta }_𝐐`$ could, in principle, be tested by neutron scattering experiments at changing external fields. On the experimental side, Bragg peaks have indeed been observed in BSCCO for $`H500`$G by \[Cubitt et al. 1993\] (?). More recently, \[Kim et al. 1999\] (?) have used the decoration technique to determine the structural properties of the vortex lattice in BSCCO and confirmed the existence of the perturbative and the random-manifold regime with $`\zeta _{\mathrm{rf}}0.22`$. On larger length scales their data exhibited non-equilibrium features such that the true asymptotic regime was not reached. So far, the resolution is, however, too weak to determine the $`\gamma `$ dependence of $`\overline{\eta }`$. Contact to the neutron scattering experiment can be made by the structure factor $`\widehat{S}(𝐤_{},k_z)`$ $`=`$ $`{\displaystyle \underset{𝐗}{}}{\displaystyle 𝑑z\overline{e^{i𝐤_{}[𝐗+𝐮(𝐗,z)𝐮(\mathrm{𝟎},0)]+ik_zz}}}`$ (6.34) $``$ $`{\displaystyle \underset{𝐗}{}}{\displaystyle 𝑑ze^{i(𝐤_{}𝐗+k_zz)}e^{\frac{1}{2}k_\alpha k_\beta W_{\alpha \beta }(𝐗,z)}},`$ where we have used the Gaussian approximation (which is correct to order $`ϵ`$) for the distribution of $`𝐮`$ and the definition (6.29) for $`W_{\alpha \beta }`$. With $`𝐤_{}=𝐐+𝐪_{}`$ and $`|𝐪_{}||𝐐|`$ we get $`\widehat{S}(𝐐+𝐪_{},k_z)\rho _0{\displaystyle d^3re^{i(𝐪_{}𝐱+k_zz)}S(𝐐,𝐫)}.`$ (6.35) $`\widehat{S}(𝐤)`$ describes the divergence of the scattered intensity with vanishing $`𝐪`$ in the vicinity of a reciprocal lattice vector $`𝐐`$. The above integral is dominated for small $`𝐪`$ by the large scale $`S(𝐐,𝐫)`$, provided $`\overline{\eta }_𝐐(\gamma )<3`$. It is thus the sub-algebraic growth of the displacements (LABEL:displace\_ab) that gives rise to the Bragg peaks, hence the name ‘Bragg glass’. In the special cases $`\gamma =0`$ and $`\gamma =1`$ one obtains $`\widehat{S}(𝐐+𝐪_{},k_z)\left(q_{}^2+{\displaystyle \frac{c_{44}}{c_{66}}}k_z^2\right)^{[3+\overline{\eta }_𝐐(\gamma )]/2}.`$ (6.36) To summarize the situation in impure bulk superconductors: it turns out that these still show a quasi-long-range ordered ‘Bragg-glass’ phase which is described by a non-universal power-like decay of the order parameter correlations. In particular, the decay-exponent $`\overline{\eta }_𝐐`$ depends on the ratio $`\gamma =\stackrel{~}{c}_{66}/\stackrel{~}{c}_{11}`$ of the elastic constants, similar to $`2D`$ pure crystals at their melting temperature. For weak disorder we find a crossover of the structural correlation functions $`S(𝐐,𝐫)`$ from a Larkin regime, where perturbation theory applies and $`S(𝐐,𝐫)`$ decays exponentially, to the random-manifold regime with a stretched exponential decay of $`S(𝐐,𝐫)`$ and eventually to the asymptotic Bragg-glass regime. In addition to the disorder-averaged positional correlation function $`S(𝐐,𝐫)`$ it is interesting to also consider the glass correlation functions $`S_{\mathrm{PG}}(𝐐,𝐫)`$ and $`C_{\mathrm{VG}}(𝐐,𝐫)`$. In the framework of the functional RG in $`d=4ϵ`$ dimensions, discussed above, it is easy to show that to first order $`ϵ`$ one finds \[Bogner et al. 2000\] $`S_{\mathrm{PG}}(𝐐,𝐫)S_{\mathrm{PG}}^{(0)}(𝐐,𝐫)\left[1+ϵ{\displaystyle \underset{m1}{}}c_m\left({\displaystyle \frac{T}{r^2}}\right)^{2m}+O(ϵ^2)\right],`$ (6.37) where $`S_{\mathrm{PG}}^{(0)}(𝐐,𝐫)=[S^{(0)}(𝐐,𝐫)]^2`$. $`S^{(0)}(𝐐,𝐫)=\mathrm{exp}[\frac{1}{2}Q_\alpha Q_\beta W_{\mathrm{th},\alpha \beta }(𝐫)]`$ denotes the pair-correlation function for TLRO of the pure system, which is finite for $`r\mathrm{}`$ if $`d>2`$ and is reduced with respect to unity merely by the standard Debye-Waller factor. For large distances $`|𝐫|`$ the leading correction in (6.37) comes from the term with $`m=1`$. Since $`c_1>0`$, the disorder increases slightly the glass order with respect to the pure system as has to be expected. With $`S_{\mathrm{PG}}(𝐐,𝐫)/[S(𝐐,𝐫)]^2\mathrm{}`$ for $`r\mathrm{}`$, we conclude that the Bragg-glass regime is indeed a positional glass. Using the relation (2.25) it is also easy to show to order $`ϵ`$ that $`C_{\mathrm{VG}}(𝐫)`$ decays exponentially in $`d<4`$ dimensions (and algebraically for $`d=4`$). Expanding all corrections from the disorder with respect to $`ϵ=4d`$ we get $`C_{\mathrm{VG}}(𝐫)e^{\frac{1}{2}(\delta \varphi (𝐫)\delta \varphi (\mathrm{𝟎}))^2_{\mathrm{th}}}\left[1+ϵ{\displaystyle \underset{m0}{}}\stackrel{~}{c}_mT^2\left({\displaystyle \frac{T}{r}}\right)^{2m}+O(ϵ^2)\right],`$ (6.38) where $`(\delta \varphi (𝐫)\delta \varphi (\mathrm{𝟎}))^2_{\mathrm{th}}=2\left({\displaystyle \frac{2\pi }{a^2}}\right)^2T{\displaystyle _𝐪}{\displaystyle \frac{q_{}^2[1\mathrm{cos}(𝐪𝐱)]}{q^4(c_{66}q_{}^2+c_{44}q_{}^2)}}{\displaystyle \frac{1}{ϵ}}|𝐱|^ϵ`$ (6.39) denotes the phase fluctuations of the order parameter in the pure system (see \[Moore 1992\] (?)). Although the leading correction ($`m=0`$) from the disorder increases $`C_{\mathrm{VG}}(𝐫)`$, it does not compensate its exponential decay originating from strong thermal fluctuations in the pure system. Whether this result remains qualitatively correct if higher order terms in $`ϵ`$ are taken into account remains an open question. For the moment we conclude that to order $`ϵ`$ there is no phase-coherent vortex-glass order in the Bragg-glass phase. Our findings are not necessarily in contradiction to the result of \[Dorsey et al. 1992\] (?) who, starting from a random-$`T_\mathrm{c}`$ GL model and using mean-field theory, found a transition to a phase with phase-coherent glass order. However, this calculation – once fluctuations are taken into account – proves the existence of a vortex glass transition with a diverging susceptibility $`\chi _{\mathrm{VG}}=d^drC_{\mathrm{VG}}(𝐫)`$ only in $`d=6ϵ`$ dimensions ($`ϵ1`$). In this region ($`d>4`$) our calculation also gives both $`S_{\mathrm{PG}}(𝐐,𝐫)`$ and $`C_{\mathrm{VG}}(𝐫)`$ non-zero for $`r\mathrm{}`$. ### 6.3 Stability of Bragg glass In this chapter we have been treating so far the vortex-line lattice in the elastic approximation, i.e., we have disregarded topological defects such as dislocations or disclinations. In the following we will consider the stability of the Bragg-glass phase with respect to dislocations. This problem was considered by several authors \[Giamarchi and Le Doussal 1995, Kierfeld et al. 1997, Gingras and Huse 1996, Carpentier et al. 1996, Ertaş and Nelson 1996, Giamarchi and Le Doussal 1997, Fisher 1997, Kierfeld 1998, Koshelev and Vinokur 1998\]. Since the disorder seen by the dislocation is already partially screened due to the elastic deformations (a fact which was overlooked in the early discussion of the elastic approximation by \[Fisher et al. 1991a\] (?)), we start this section with a brief discussion of the effective disorder strength $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}(R)`$ acting on scale $`R`$ in the various regimes. We will thereby ignore all numerical coefficients of order unity and restrict ourselves to the case $`d=3`$. For simplicity, we restrict ourselves here to the case $`c_{11}c_{66}`$ such that we can assume $`\mathbf{}_{}𝐮=0`$. It is then convenient to rewrite the Hamiltonian using the coordinate $`z_t=(c_{66}/c_{44})^{1/2}z`$ introduced in equation (6.10). The elastic part of the Hamiltonian is now isotropic with $`c_{44}`$ and $`c_{66}`$ replaced by their geometric mean $`(c_{44}c_{66})^{1/2}`$. The same transformation in the pinning part changes $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ into $`(c_{44}/c_{66})^{1/2}\stackrel{~}{\mathrm{\Delta }}(𝐮)`$. It is also convenient to introduce the (anisotropic) distance $`R_t`$, $`R_t^2=𝐱^2+z_t^2.`$ (6.40) $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R)`$ was introduced already in equation (6.28). We use here (6.27) to find explicit expressions for the mean square displacement $`\overline{𝐮^2}(R_t)`$. To obtain an estimate for $`\overline{𝐮^2}(R_t)W(R_t)/2`$ on the scale $`R_t`$ one equates the elastic energy $`E_{\mathrm{el}}(c_{44}c_{66})^{1/2}\overline{𝐮^2}R_t`$, with the fluctuation of the pinning energy $`E_{\mathrm{pin}}\left({\displaystyle \frac{c_{44}}{c_{66}}}\right)^{1/4}R_t^{3/2}\left(\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)\overline{𝐮^2}\right)^{1/2}.`$ (6.41) $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)`$ denotes the curvature of the effective potential correlator at $`𝐮=0`$ on the length scale $`R_t`$. The estimate for the pinning energy corresponds to the use of perturbation theory applied on the renormalized random potential. In this way we obtain $`\overline{𝐮^2}c_{44}^{1/2}c_{66}^{3/2}\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)R_t=\xi ^2{\displaystyle \frac{\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)}{\stackrel{~}{\mathrm{\Delta }}^{(2)}}}{\displaystyle \frac{R_t}{R_\xi }},`$ (6.42) where $`\stackrel{~}{\mathrm{\Delta }}^{(2)}\stackrel{~}{\mathrm{\Delta }}^{(2)}(R_\xi )`$ denotes the bare value of $`\stackrel{~}{\mathrm{\Delta }}^{(2)}(R_t)`$. Now we consider the different regimes already discussed in section 6.2. (i) In the perturbative regime $`R_t<R_\xi \xi ^2c_{44}^{1/2}c_{66}^{3/2}/\stackrel{~}{\mathrm{\Delta }}^{(2)}`$, $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}\stackrel{~}{\mathrm{\Delta }}^{(2)}`$ in equation (6.42). (ii) In the manifold region, $`R_\xi <R_t<R_a`$, where $`\overline{𝐮^2}\xi ^2(R_t/R_\xi )^{2\zeta _{\mathrm{rm}}}`$, we find from equation (6.42) $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)\stackrel{~}{\mathrm{\Delta }}^{(2)}\left({\displaystyle \frac{R_t}{R_\xi }}\right)^{2\zeta _{\mathrm{rm}}1}.`$ (6.43) (iii) Finally, in the asymptotic Bragg-glass region, $`R_a<R_t`$, we get from $`\overline{𝐮^2}a^2\mathrm{ln}(R_t/R_a)`$ $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)\stackrel{~}{\mathrm{\Delta }}^{(2)}{\displaystyle \frac{R_\xi }{R_t}}{\displaystyle \frac{a^2}{\xi ^2}}{\displaystyle \frac{a^2c_{44}^{1/2}c_{66}^{3/2}}{R_t}}.`$ (6.44) Thus, $`\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}(R_t)`$ is independent of the bare disorder strength. Here we have to ignore the logarithm since it originates from the summation over many different length scales. We use now these expressions to obtain an estimate for the energy of a dislocation loop. For simplicity, we assume an isotropic (i.e., almost circular) loop of linear size $`R_t`$. In the pure system, its elastic energy is $`E_{\mathrm{el}}^{\mathrm{loop}}b^2(c_{44}c_{66})^{1/2}R_t\mathrm{ln}(R_t/a_1),`$ (6.45) where $`|𝐛|=a_{\mathrm{}}`$ and $`𝐛`$ denotes the Burgers vector of the dislocation. The logarithm results from the fact that in a distance $`R_t`$ from the dislocation centre the displacement $`𝐮_{\mathrm{disloc}}`$ produced by the dislocation line obeys $`|\mathbf{}𝐮_{\mathrm{disloc}}|\frac{|𝐛|}{2\pi R_t}`$. Integration over the plane perpendicular to the dislocation line then yields $`a_{\mathrm{}}^2(c_{44}c_{66})^{1/2}\mathrm{ln}(R_t/a_1)`$ for the energy per unit length, where $`a_1`$ acts as a small distance cut-off. In a random system, the situation is more complicated. Since the displacement $`𝐮_{\mathrm{disloc}}`$ created by a dislocation line obeys the saddle-point equation $`\delta /\delta 𝐮=0`$ as well as $`𝑑𝐮=𝐛`$, the elastic energy of the dislocation now depends also on the disorder. Since the situation apparently is rather involved, we will adopt here a simplified picture using the following observation: The displacement in the neighbourhood of a dislocation has two sources: Its very existence leads in the distance $`R_t`$ from the dislocation to a (essentially tangential) displacement of order $`a`$. The disorder, on the other hand, creates displacements (both in the absence or in the presence of a dislocation) of order $`a`$ or larger only on scales $`R_t>R_a`$. In calculating the energy of a loop of linear size $`R_t>R_a`$ in the disordered case we estimate the elastic energy of the dislocation loop therefore by $`E_{\mathrm{el}}^{\mathrm{loop}}a^2(c_{44}c_{66})^{1/2}R_t\mathrm{ln}(R_a/a_1),`$ (6.46) since distortions originating from the random potential are small compared to $`a`$ on scales $`RR_a`$ but are dominating on larger scales. So far we have ignored the dispersion of the elastic constant $`c_{44}`$, which becomes relevant on scales small compared to $`\lambda `$. On these scales the lattice is softer by a factor $`(R_t/\lambda )^2`$. In order to take this effect into account, we replace the small scale cut-off $`a_1`$ in (6.46) by $`(a_1^2+\lambda ^2)^{1/2}`$. The distortions on scales larger than $`R_a`$ are dominated by the disorder and lead to an energy gain of the order (again omitting the factor $`\mathrm{ln}(R_t)`$ in $`\overline{𝐮^2}`$, since it comes from the summation over displacements on different length scales) $`E_{\mathrm{pin}}^{\mathrm{loop}}\left({\displaystyle \frac{c_{44}}{c_{66}}}\right)^{1/4}R_t^{3/2}\left(\stackrel{~}{\mathrm{\Delta }}_{\mathrm{eff}}^{(2)}\overline{𝐮^2}\right)^{1/2}a^2(c_{44}c_{66})^{1/2}R_t`$ (6.47) where we used equations (6.41) and (6.44). The total loop energy of a dislocation loop can therefore be written as $`E^{\mathrm{loop}}a^2(c_{44}c_{66})^{1/2}R_t\left(\mathrm{ln}\left({\displaystyle \frac{R_a^2}{a_1^2+\lambda ^2}}\right)c_1\right),`$ (6.48) where $`c_1`$ is a constant of order unity. (It is easy to see by analogous arguments, but using equations (6.42) or (6.43) for $`\mathrm{\Delta }_{\mathrm{eff}}(R_t)`$, that dislocation loops of size $`R_t<R_a`$ always have a positive energy.) From (6.48) we conclude that the system is stable against the formation of a dislocation loop as long as $`R_ac_2(a_1^2+\lambda ^2)^{1/2}`$ (6.49) with $`c_2e^{c_1}`$. As shown by \[Kierfeld 1998\] (?) for strong disorder (i.e., $`R_\xi a`$), this relation can be rewritten in the form of a Lindemann criterion. Indeed, in the manifold regime (note that the Larkin regime vanishes for strong disorder), where $`{\displaystyle \frac{\overline{𝐮^2}(R_a)}{\overline{𝐮^2}(a)}}\left({\displaystyle \frac{R_a^2}{a^2+\lambda ^2}}\right)^{\zeta _{\mathrm{rm}}}`$ (6.50) and $`\overline{𝐮^2}(R_a)a^2`$ one obtains from (6.49) $`\overline{[𝐮(a)𝐮(0)]^2}^{1/2}c_{\mathrm{Li}}a,`$ (6.51) which is the Lindemann criterion appropriate for disordered systems in which the mean-square displacement of a single vortex line diverges. $`c_{\mathrm{Li}}=c_2^{\zeta _{\mathrm{rm}}}`$ denotes the Lindemann number. The main conclusion from this consideration is that the elastic, dislocation-free Bragg glass is stable as long as the criterion (6.49) or, for strong disorder, equation (6.51)\] is fulfilled. A few concluding remarks are in order: * The criterion (6.51) \[or, equivalently, equation (6.49)\] was also derived by \[Kierfeld et al. 1997\] (?) and \[Carpentier et al. 1996\] (?) via a variational treatment for a layered system with the magnetic field parallel to the layers. In this model only dislocations with Burgers vectors parallel to the layers are allowed. The most important result is the determination of the Lindemann number $`c_{\mathrm{Li}}0.14`$. We refer the reader to these papers for further details. * \[Ertaş and Nelson 1996\] (?) took equation (6.51) as a starting point for a discussion of the onset of irreversibility and entanglement of vortex lines in high-$`T_c`$ materials. * \[Kierfeld 1998\] (?), on the basis of equation (6.51), and including the dispersion of the elastic constants in detail, calculated the stability boundaries of the Bragg-glass phase for YBCO and BSCCO. The resulting phase diagram is depicted schematically in figure 11. Since the shear modulus decreases for large and small fields, respectively, there are two corresponding stability boundaries. * \[Fisher 1997\] (?) undertook a much more detailed study of the stability of a defect loop in the random-field $`XY`$ model which corresponds to the layered model mentioned in (i). Although the details of his energy estimate for the dislocation loop are slightly different from those presented here (using a statistical tilt symmetry, he estimates $`E_{\mathrm{el}}^{\mathrm{loop}}`$ by equation (6.45) but includes also rare fluctuations in the energy gain from the disorder, which also include logarithmic corrections), his final conclusions concerning the stability of the defect-free phase are essentially identical to those presented here. * \[Ryu et al. 1996b\] (?) numerically investigated the field-driven transition from a dislocation-free to a dislocation-dominated phase in an impure layered superconductor and found good quantitative agreement with the estimates made above as well as with the neutron-scattering data of \[Cubitt et al. 1993\] (?), who found the disappearance of Bragg peaks in BSCCO for $`H`$500G. * Most recently, \[Kierfeld and Vinokur 1999\] (?) used a random-stress model of the form (3.33) as an effective model to describe dislocation lines in an impure superconductor. The correlations of the random-stress field $`𝝁(𝐫)`$ on different length scales are chosen in such a way that the correct roughness exponents in the manifold and the Bragg-glass regime, respectively, are reproduced. Considering in particular dislocation lines with main orientation parallel to the magnetic field, which enter and leave the sample at its surface, \[Kierfeld and Vinokur 1999\] (?) found a global phase diagram, which includes – besides the elastic Bragg-glass phase with vanishing dislocation density – an amorphous vortex-glass and a vortex-liquid phase separated by a first-order transition line which terminates in a critical point (see figure 11). ## 7 Vortices driven far from equilibrium As the main result of the previous chapter we found that weak pinning reduces the positional order of the vortex lattice in a bulk superconductor from long-range order to quasi-long-range order. The discussion was restricted to thermodynamic equilibrium. However, since the electric resistivity is one of the most relevant physical properties of superconductors, it is important to study the driven non-equilibrium situation. Since the efficiency of pinning is related to the structure of the vortex system, it is of particular interest to characterize this structure. In addition, since the melting transition of the vortex lattice shows up as a pronounced shoulder in the transport characteristic (see, e.g., \[Safar et al. 1992\] (?), \[Kwok et al. 1992\] (?)), it is desirable to locate this transition in the non-equilibrium situation. The VLL shows the three regimes of creep, depinning, and flow in its transport characteristic, which were already described for manifolds in chapter 3.2 (cf. figure 5). We recall that in a superconductor the vortex velocity is proportional to the electric field and the driving force is proportional to the electric current density. So far, we have mainly focused on the creep regime when we addressed transport properties in the previous chapters. In this regime, which is close to equilibrium, the dynamical response is determined within the scaling approach by the structure of the VLL in equilibrium. In particular, the logarithmic roughness of the elastic vortex glass (described by a roughness exponent $`\zeta =0`$) results in a drift velocity $`v(F)(F/\eta _0)e^{U(F)/T},`$ (7.1) where the effective barrier height $`U(F)`$ scales with the driving force $`F`$ according to $`U(F)F^\mu `$ (7.2) with the creep exponent \[Nattermann 1990\] $`\mu ={\displaystyle \frac{d2}{2}}`$ (7.3) \[cf. equations (3.71) to (3.73) with the appropriate substitution for the dimension of the manifold\]. In $`d<2`$ one thus expects $`\mu <0`$, which means that the system has a linear transport characteristic at small driving forces. Only in $`d2`$ we can expect to find true superconductor with a vanishing linear resistivity. The ($`d=2`$)-dimensional case is marginal since it has the creep exponent $`\mu =0`$. There the effective barrier height depends logarithmically on the driving force and results in a power-law transport characteristic as found in equations (4.56) and (5.30) for the superconducting film in parallel and perpendicular field. In this chapter we will focus on the flow regime, which is far from equilibrium but accessible to a theoretical analysis because it is, roughly speaking, close to the pure case since disorder is dynamically ‘averaged out’. We will elucidate this point later on. Although the flow regime might appear to be of no special interest at first sight, only recently it turned out to be quite non-trivial, including phenomena such as non-equilibrium phase transitions. To perform the theoretical analysis of a stationary state at average velocity $`𝐯`$ it is convenient to define the displacement $`𝐮(𝐗,𝐳,t)𝐱(𝐗,𝐳,t)𝐗𝐯t`$ by subtracting the average temporal displacement, such that $`𝐮`$ still can be considered as a small quantity. $`𝐗`$ is then the ideal vortex position measured in a frame moving with velocity $`𝐯`$. In analogy to manifolds, the equation of motion for the vortex lattice, $`\eta _0\dot{𝐮}=𝐅^{\mathrm{int}}+𝐅^{\mathrm{pin}}+𝐅\eta _0𝐯+𝜻,`$ (7.4) is over-damped. In comparison to equation (3.53) for a single vortex line, equation (7.4) includes the vortex density $`\rho _0`$ as additional factor, such that the latter equation is an equation for the force density in the $`d`$-dimensional space. Accordingly, the left-hand side is the Bardeen-Stephen friction force $`\eta _0\dot{𝐮}`$ \[Bardeen and Stephen 1965\], with the friction coefficient $`\eta _0BH_{\mathrm{c2}}/\rho _\mathrm{n}c^2`$ referring to the force per unit volume, whereas the previous expression (3.50) refers to the force per unit length of a vortex line. $`𝐅^{\mathrm{int}}`$ is the force acting on a vortex due to its interaction with other vortices. In a harmonic approximation for the elastic phase one has $`𝐅^{\mathrm{int}}𝐅^{\mathrm{el}}`$. The instantaneous elastic force $`𝐅^{\mathrm{el}}`$ depends linearly on $`𝐮`$ and reads $`F_\alpha ^{\mathrm{el}}(𝐪,t)`$ $`=`$ $`\mathrm{\Gamma }_{\alpha \beta }(𝐪)u_\beta (𝐪,t),`$ (7.5a) $`\mathrm{\Gamma }_{\alpha \beta }(𝐪)`$ $`=`$ $`{\displaystyle \underset{p}{}}[c_pq_{}^2+c_{44}q_{}^2]P_{\alpha \beta }^p(𝐪_{}),`$ (7.5b) where $`\mathrm{\Gamma }`$ represents the elastic dispersion, $`p=L,T`$ stands for the longitudinal/transverse polarization, $`c_pc_{11},c_{66}`$ respectively, and $`q_{}^2=q_x^2+q_y^2`$. As far as we consider a general dimension $`d`$, $`𝐱`$ and $`𝐪_{}`$ are two-dimensional and $`𝐳`$ and $`𝐪_{}`$ are $`(d2)`$-dimensional. $`𝐅^{\mathrm{pin}}`$ is the pinning force density $`𝐅^{\mathrm{pin}}(𝐗,𝐳,t)=\rho _0\mathbf{}_{}V(𝐗+𝐯t+𝐮(𝐗,𝐳,t),𝐳),`$ (7.6) which implicitly depends on the displacement $`𝐮`$. The pinning potential $`V`$ and the thermal noise $`𝜻`$ are assumed to be Gaussian distributed with correlators (6.2) and $`\zeta _\alpha (𝐗,𝐳,t)\zeta _\beta (\mathrm{𝟎},\mathrm{𝟎},0)=2\eta _0T\rho _0\delta _{𝐗,\mathrm{𝟎}}\delta _{\alpha \beta }\delta (𝐳)\delta (t)`$ (7.7) in analogy to equation (3.52). ### 7.1 Qualitative aspects On the basis of this equation of motion we can become more specific in what sense disorder is ‘dynamically averaged out’ in the limit of large drift velocities. In this limit the pinning force $`𝐅^{\mathrm{pin}}(𝐗,𝐳,t)`$ acting on a fixed vortex element at position $`(𝐗,𝐳)`$ in the comoving frame changes rapidly as a function of time. Since the VLL has finite response times on finite length scales, the displacements induced by the pinning force decrease with increasing $`𝐯`$ and the pinning potential is effectively averaged (‘washed’) out. Therefore the effect of disorder vanishes in the limit of very large driving force. To substantiate this relation it is instructive to consider a single vortex line $`𝐗`$ moving with strictly constant velocity (i.e., $`𝐱=𝐗+𝐯t`$), neglecting its response to pinning. In this approximation the vortex line experiences a time dependent force $`𝐅^{\mathrm{pin}}(𝐗,𝐳,t)=\rho _0\mathbf{}_{}V(𝐗+𝐯t,𝐳)`$. To some extent the effect of this force can be compared to an additional thermal noise with a certain ‘shaking temperature’ $`T^{\mathrm{sh}}`$. We choose the $`𝐱(x,y)`$ coordinates such that $`𝐯=(v,0)`$ points into the direction of the first basis vector $`𝐞_x`$. Then $`T_\alpha ^{\mathrm{sh}}`$ $``$ $`{\displaystyle \frac{1}{2\eta _0\rho _0}}{\displaystyle 𝑑td^{d2}z\overline{F_\alpha ^{\mathrm{pin}}(𝐗,𝐳,t)F_\alpha ^{\mathrm{pin}}(𝐗,\mathrm{𝟎},0)}}`$ (7.8a) $`=`$ $`{\displaystyle \frac{\rho _0}{2\eta _0v}}{\displaystyle 𝑑x_1_\alpha ^2\mathrm{\Delta }(x_1𝐞_x)},`$ $`T_x^{\mathrm{sh}}`$ $`=`$ $`0,`$ (7.8b) $`T_y^{\mathrm{sh}}`$ $``$ $`{\displaystyle \frac{\rho _0\mathrm{\Delta }_0}{2\eta _0v\xi ^3}},`$ (7.8c) where $`\alpha =x,y`$ is not to be summed over implicitly. These equations are strictly analogous to the equations (LABEL:T.sh.mani) for manifolds. The concept of the ‘shaking temperature’ was introduced by \[Koshelev and Vinokur 1994\] (?), who considered in particular the case $`d=2`$ and introduced the ‘incoherent’ shaking temperature $`T^{\mathrm{sh}}\frac{1}{2}_\alpha T_\alpha ^{\mathrm{sh}}`$. For qualitative purposes they considered the system driven through disorder as being subject to a total effective temperature $`T_{\mathrm{eff}}(v)=T+T^{\mathrm{sh}}(v)`$. Then the vortex lattice can be expected to melt at a velocity-dependent temperature $`T_\mathrm{m}(v)=T_\mathrm{m}T^{\mathrm{sh}}(v),`$ (7.9) where $`T_\mathrm{m}T_\mathrm{m}(v=\mathrm{})`$ is the melting temperature of the pure system. Equivalently, the inverted function $`v_\mathrm{m}(T)`$ defines a melting velocity above which the vortices freeze into a solid (‘dynamic freezing’). $`T^{\mathrm{sh}}`$ vanishes for increasing $`v`$ and hence both $`v_\mathrm{m}`$ and the corresponding driving force $`F_\mathrm{m}=F(v_\mathrm{m})`$ increase with increasing $`T`$ for fixed pinning strength. From the velocity dependence of $`T_y^{\mathrm{sh}}`$ one expects $`T_\mathrm{m}T_\mathrm{m}(v){\displaystyle \frac{1}{v}}{\displaystyle \frac{1}{F}}`$ (7.10) at large driving forces. This relation is consistent with experimental observations \[Bhattacharya and Higgins 1993, Hellerqvist et al. 1996\] and numerical simulations \[Koshelev and Vinokur 1994\]. A quantitative analysis of the pinning effects requires taking into account the interaction between the vortex lines. For this purpose \[Koshelev and Vinokur 1994\] (?) introduced a ‘coherent’ shaking temperature, which they found to scale like $`T_{\mathrm{coh}}^{\mathrm{sh}}(v)(v_{\mathrm{rel}}/v)T^{\mathrm{sh}}(v)`$ with a characteristic velocity scale $`v_{\mathrm{rel}}`$, taking into account only the transverse response of the VLL in $`d=2`$. However, the question of what physical properties of the system can be described by such an effective temperature is quite subtle. This is evident from equation (LABEL:intro.Tsh), which shows that the shaking effect is anisotropic and that one should distinguish the direction parallel to the driving force from perpendicular directions. A more careful treatment of disorder will be presented below to evaluate the effects of disorder on a solid VLL in an elastic approximation (section 7.2) and to characterize the structure of the driven lattice, before one can analyze the stability of the VLL with respect to the proliferation of topological defects, from what one can locate the melting transition (section 7.3). ### 7.2 Moving lattice To characterize the structure of the vortex lattice driven in disorder, we start from the assumption of an elastic and topologically ordered phase, as exists in the absence of disorder and for low temperatures in $`d2`$. The first step is to treat pinning perturbatively in the high-velocity regime. This analysis is most convenient in Fourier representation, which reads for the displacement $`𝐮(𝐫,t)={\displaystyle _\omega }{\displaystyle _𝐪}e^{i(𝐪𝐫\omega t)}𝐮(𝐪,\omega ),`$ (7.11) where $`_\omega =𝑑\omega /2\pi `$ and $`_𝐪=d^dq/(2\pi )^d`$ is restricted to the first Brillouin zone. The dynamic response function $`G`$ of the pure system is ($`p=L,T`$) $`G_{\alpha \beta }(𝐪,\omega )`$ $`=`$ $`{\displaystyle \underset{p}{}}G^p(𝐪,\omega )P_{\alpha \beta }^p(𝐪),`$ (7.12a) $`G^p(𝐪,\omega )`$ $`=`$ $`[i\eta _0\omega +c_pq_{}^2+c_{44}q_{}^2]^1.`$ (7.12b) To zeroth order in $`𝐮`$ the pinning force acting on the VLL in the comoving frame has the correlation $`\mathrm{\Xi }_{\alpha \beta }(𝐗,𝐳,t)`$ $``$ $`\overline{F_\alpha ^{\mathrm{pin}}(𝐗,𝐳,t)F_\beta ^{\mathrm{pin}}(\mathrm{𝟎},\mathrm{𝟎},0)}`$ (7.13a) $`=`$ $`\rho _0^2_\alpha _\beta \mathrm{\Delta }(𝐗+𝐯t)\delta (𝐳),`$ $`\mathrm{\Xi }_{\alpha \beta }(𝐪,\omega )`$ $`=`$ $`\rho _0^2{\displaystyle \underset{𝐐}{}}k_\alpha k_\beta \mathrm{\Delta }(𝐤)\delta (\omega +𝐯𝐤)`$ (7.13b) $`=`$ $`\rho _0^2{\displaystyle \underset{𝐐}{}}\mathrm{\Delta }_{\alpha \beta }^{(2)}(𝐤)\delta (\omega +𝐯𝐤),`$ where $`𝐐`$ is a reciprocal lattice vector (RLV) and $`\mathrm{\Delta }_{\alpha \beta }^{(2)}(𝐤)k_\alpha k_\beta \mathrm{\Delta }(𝐤)`$. The wave vector $`𝐤𝐐+𝐪`$ covers the whole Fourier space without restriction to the first Brillouin zone (in contrast to $`𝐪`$). In the following we assume that the vortex lattice is oriented with one principal axis parallel to the average velocity. \[Schmid and Hauger 1973\] (?) argued that the vortex lattice orients itself in this direction because this is the direction of minimum entropy production and of least power dissipation. One can arrive at the same conclusion from a stability analysis \[Müllers and Schmid 1995\]. In this case there are RLVs $`𝐐`$ which are perpendicular to the velocity $`𝐯`$ and which play an important role for the dynamics. An inspection of the correlator (LABEL:intro.Xi) for $`\omega =0`$ and $`𝐪=\mathrm{𝟎}`$ shows that the disorder correlator evaluated at these RLVs will determine the lattice distortions on large length and time scales. On these scales the system can be described well by approximating $`\mathrm{\Xi }_{\alpha \beta }(𝐪,\omega )\mathrm{\Xi }_{\alpha \beta }\delta (\omega +𝐯𝐪)`$ (7.14) with $`\mathrm{\Xi }_{\alpha \beta }=\rho _0^2{\displaystyle \underset{𝐐(𝐯)}{}}\mathrm{\Delta }_{\alpha \beta }^{(2)}(𝐐)\stackrel{~}{\mathrm{\Delta }}^{(2)}\delta _{\alpha y}\delta _{\beta y}`$ (7.15) and $`\stackrel{~}{\mathrm{\Delta }}^{(2)}`$ as given in equation (6.25). From the pinning force correlator (LABEL:intro.Xi) the zero-temperature displacement correlations follow via $`W_{\alpha \beta }(𝐫,t)`$ $``$ $`\overline{[u_\alpha (𝐫,t)u_\alpha (\mathrm{𝟎},0)][u_\beta (𝐫,t)u_\beta (\mathrm{𝟎},0)]}`$ (7.16) $`=`$ $`2{\displaystyle _\omega }{\displaystyle _𝐪}(1e^{i(𝐪𝐫\omega t)})G_{\alpha \gamma }(𝐪,\omega )\mathrm{\Xi }_{\gamma \delta }(𝐪,\omega )G_{\delta \beta }(𝐪,\omega ).`$ Within this lowest-order perturbative approach one finds explicitly on large scales $`|𝐫|\mathrm{}`$ that the fluctuations of $`u_x`$ are finite (at $`T=0`$), whereas $`u_y`$ is rough in dimensions $`d3`$ \[Balents and Fisher 1995, Giamarchi and Le Doussal 1996, Balents et al. 1997\]: $`W_{xx}(𝐫,t=0)`$ $``$ $`{\displaystyle \frac{2\rho _0\stackrel{~}{\mathrm{\Delta }}(\mathrm{𝟎})}{\eta _0^2v^2}}\left({\displaystyle \frac{2\pi }{\xi }}\right)^{d2},`$ (7.17a) $`W_{yy}(𝐫,t=0)`$ $``$ $`{\displaystyle \frac{a^{3d}\mathrm{\Xi }_{yy}}{\eta _0vc_{\mathrm{eff}}}}𝒲\left(\rho _0\mathrm{max}\{{\displaystyle \frac{c_{11}|x|}{\eta _0v}},y^2,{\displaystyle \frac{c_{11}}{c_{44}}}𝐳^2\}\right),`$ (7.17b) with an effective elastic constant $`c_{\mathrm{eff}}=c_{11}`$ in $`d=2`$ and $`c_{\mathrm{eff}}=(c_{11}c_{44})^{1/2}`$ in $`d=3`$ (in general dimension, $`c_{\mathrm{eff}}=c_{11}^{2d/2}c_{44}^{d/21}`$) and a scaling function with $`𝒲(0)=0`$ and $`𝒲(s)`$ $``$ $`\{\begin{array}{cc}\text{const.},\hfill & d>3,\hfill \\ \mathrm{ln}(s),\hfill & d=3,\hfill \\ s^{(3d)/2},\hfill & d<3,\hfill \end{array}`$ (7.21) for $`s\mathrm{}`$. In comparison to the equilibrium case, the driven vortex lattice is roughened by disorder only in dimensions $`d3`$. Thus, in (the somewhat academic) dimensions $`3<d4`$, the roughness disappears due to the drive, and disorder is washed out substantially by driving the VLL. It is further remarkable that the roughness of the driven lattice (as found in lowest order perturbation theory) arises from compression modes. In contrast to this, in equilibrium even an incompressible lattice is roughened. Although the shaking temperature (LABEL:intro.Tsh) suggests larger fluctuations of $`u_y`$ than of $`u_x`$ in agreement with equation (LABEL:W.pert.drive), the roughness of the lattice cannot be described by adding the shaking temperature (LABEL:intro.Tsh) to the physical temperature of the system. As long as this temperature is finite, one would expect the lattice to be flat in $`d>2`$, which is inconsistent with the roughness found above in $`2<d3`$. In analogy to the equilibrium case we may define a dynamic Larkin length, beyond which the perturbative approach breaks down, from $`W(𝐫,t=0)W_{xx}(𝐫,t=0)+W_{yy}(𝐫,t=0)=\xi ^2`$, to which the $`y`$ component of the displacements gives the dominant contribution according to equation (LABEL:W.pert.drive). Because of the anisotropy of $`W_{yy}(𝐫,t=0)`$, the Larkin length strongly depends on the orientation of $`𝐫`$: $`L_\xi ^{(y)}`$ $``$ $`\{\begin{array}{cc}a\mathrm{exp}\left(\frac{\eta _0v\xi ^2\sqrt{c_{11}c_{44}}}{\mathrm{\Xi }_{yy}}\right),\hfill & d=3,\hfill \\ a\left(\frac{\eta _0v\xi ^2c_{\mathrm{eff}}}{\mathrm{\Xi }_{yy}a^{3d}}\right)^{1/(3d)},\hfill & d<3,\hfill \end{array}`$ (7.22c) $`L_\xi ^{(x)}`$ $``$ $`{\displaystyle \frac{\eta _0v}{c_{11}}}\left(L_\xi ^{(y)}\right)^2,`$ (7.22d) $`L_\xi ^{(z)}`$ $``$ $`{\displaystyle \frac{c_{44}}{c_{11}}}L_\xi ^{(y)}`$ (7.22e) \[Giamarchi and Le Doussal 1996, Balents et al. 1998a, Le Doussal and Giamarchi 1998b\]. The roughness of the driven vortex lattice (more precisely, of $`u_y`$) in $`d3`$ implies that the perturbation theory breaks down on large length scales and progress can be made only by means of a renormalization-group analysis. This analysis is very involved \[Balents et al. 1998a, Le Doussal and Giamarchi 1998b, Scheidl and Vinokur 1998b\] and only its main results will be presented. One main feature is that the disorder correlator gets renormalized in a qualitative way. While the original pinning-force correlator $`\mathrm{\Delta }_{\alpha \beta }^{(2)}(𝐤)k_\alpha k_\beta \mathrm{\Delta }(𝐤)`$ entering equation (LABEL:intro.Xi) is derived from a random potential such that $`\mathrm{\Delta }_{\alpha \beta }^{(2)}(𝐤)|_{𝐤=\mathrm{𝟎}}=0`$, under renormalization it develops a random-force character such that $`\mathrm{\Delta }_{\mathrm{},\alpha \beta }^{(2)}(𝐤)|_{𝐤=\mathrm{𝟎}}0`$. Thereby the large-scale value of the pinning-force correlator will be increased to values $`\mathrm{\Xi }_{\mathrm{},yy}>\stackrel{~}{\mathrm{\Delta }}^{(2)}`$ and, most important, $`\mathrm{\Xi }_{\mathrm{},xx}>0`$. \[Balents et al. 1997\] (?) pointed out that a roughness of the displacement component $`u_x`$ will be generated as a consequence of the finiteness of $`\mathrm{\Xi }_{\mathrm{},xx}`$. Thus, on large scales the displacement correlation function is not described by the perturbative result (7.17a) but by equation (7.17b) after the substitutions $`c_{11}c_{66}`$ and $`\mathrm{\Xi }_{yy}\mathrm{\Xi }_{\mathrm{},xx}`$. Since $`c_{66}c_{11}`$ one expects the fluctuations of $`u_y`$ to be larger than the fluctuations of $`u_x`$ on very large scales \[Balents et al. 1997, Balents et al. 1998a\]. \[Giamarchi and Le Doussal 1996\] (?) pointed out that the displacement component $`u_y`$ actually shows a frozen pattern in the laboratory frame. This means that the vortices move along ‘static channels’, see figure 12. These channels are oriented parallel to the average velocity without crossing each other, although they are rough in $`d3`$. Thus, in elastic approximation, the driven vortex lattice looks like a glass with respect to the structure, and hence one may call it ‘moving glass’ \[Giamarchi and Le Doussal 1996\]. Nevertheless, unlike the Bragg glass the ‘moving glass’ does not match all criteria of glassiness as specified in section 3.3. Since $`u_y`$ is quenched, one might expect to find a sub-linear transverse response, $`\eta _{yy}(𝐯)=dF_y(𝐯)/dv_y|_{v_y=0}=\mathrm{}`$ at finite temperature. This is, however, not the case: a finite transverse force $`F_y`$ induces a linear transverse velocity $`v_y`$ \[Le Doussal and Giamarchi 1998b, Scheidl and Vinokur 1998b\]. Only at zero temperature there is a finite critical transverse force. But this is true for any potential and not indicative for the actual relevance of disorder for the dynamics. The renormalization-group analysis reflects characteristic features of non-equilibrium, including the generation of Kardar-Parisi-Zhang (KPZ) nonlinearities, and an anisotropic renormalization of the elastic dispersion of the lattice, of the linear mobility and of the temperature \[Balents et al. 1998a, Le Doussal and Giamarchi 1998b, Scheidl and Vinokur 1998b\]. The detailed discussion of these aspects is beyond the scope of this article. ### 7.3 Moving smectic and dynamic melting In the previous section we discussed the moving vortex system in the elastic approximation and found that is has long-range order in $`d>3`$, quasi-long-range order in $`d=3`$ and only short-range order in $`d<3`$. In order to examine whether this elastic system is actually stable with respect to the generation of free dislocations, it is in principle necessary to examine their dynamic generation process, since in the non-equilibrium situation the lattice stability can no longer be examined using energetic criteria. Nevertheless, if we naively carry over our findings from equilibrium to non-equilibrium, we might expect the elastic system (the moving lattice, which is topologically ordered and where vortices move coherently) in $`d3`$ to be stable since it has (quasi-) long-range order, whereas the short-range order in $`d<3`$ should imply the instability. One can arrive at the same conclusion on the basis of a dynamical scaling analysis which was performed by \[Balents and Fisher 1995\] (?) for CDW systems. \[Balents et al. 1997\] (?) argued that because of the dominance of fluctuations of $`u_x`$ over fluctuations of $`u_y`$ on largest scales, the instability should be generated by dislocations with Burgers vectors parallel to $`𝐯`$. Due to the presence of these dislocations the channels would be decoupled dynamically, i.e. vortices in different channels could move with locally different velocities. An analogous decoupling of charge-density waves in a layered model was found by \[Vinokur and Nattermann 1997\] (?) from a variational calculation. Despite the dynamic decoupling it is possible that a static channel structure still persists. This phase is called ‘moving smectic’ \[Balents et al. 1997\] or ‘moving transverse glass’ \[Le Doussal and Giamarchi 1998b\]. The vortex density can still be modulated with quasi-long-range order in $`d=3`$. In $`d<3`$ it is possible that even this channel structure is destroyed, since the power-law roughness of $`u_y`$ may also induce dislocations with Burgers vectors having a component perpendicular to $`𝐯`$. Then the vortex system would be essentially a vortex liquid, which still has a certain anisotropy since the driving force breaks the rotation symmetry in the $`(x,y)`$ plane. Although the moving lattice can exist at sufficiently large drift velocities in $`d=3`$, the effective strength of disorder becomes larger as the velocity is reduced and therefore the moving lattice can decay first into a moving smectic and then into a moving liquid. The schematic phase diagrams \[Balents et al. 1998a, Le Doussal and Giamarchi 1998b, Scheidl and Vinokur 1998b\] for $`d=2`$ and $`d=3`$ are illustrated in figure 13. A theoretical analysis of the large-scale properties of the driven vortex system is hampered by the anisotropy of the system, the relevance of disorder and the dynamic non-linearities such as KPZ terms that govern the vortex dynamics on large scales. For these reasons a rigorous analysis has not been achieved so far. The conspiracy of these influences may actually lead to further phase transitions, such as first-order roughening transitions, which were found by \[Chen et al. 1996\] (?) for charge-density-wave systems in $`d=1,2`$. There is experimental evidence not only for dynamic melting, but also the structure of the moving lattice and moving smectic have been characterized. The transition was detected by resistive measurements \[Bhattacharya and Higgins 1993, Hellerqvist et al. 1996\] and the structure of the dynamic phases was observed by dynamic decoration techniques \[Marchevsky et al. 1997, Pardo et al. 1998, Troyanovskii et al. 1999\]. Numerical evidence for dynamic melting was found in $`d=2`$ \[Koshelev and Vinokur 1994, Ryu et al. 1996a, Faleski et al. 1996\] and also $`d=3`$ \[Domínguez et al. 1997\]. In $`d=2`$ melting can show up only as a crossover which, however, may be very sharp for weak disorder and restricted system sizes. \[Spencer and Jensen 1997\] (?) presented numerical evidence for the absence of true topological order even at large velocity in $`d=2`$. The instability of the driven two-dimensional system was demonstrated analytically by \[Aranson et al. 1998\] (?) who explicitly examined the dynamics of single dislocations and their interaction within a driven vortex lattice. They were able to show that already the presence of the KPZ terms leads to a screening of the dislocation interaction on a finite length scale (which, however, increases exponentially with the drift velocity). Consequently, dislocation–anti-dislocation pairs will unbind under the additional influence of thermal fluctuations and even more due to the shaking effect of disorder. In the remainder of this section we briefly readdress the melting transition and sketch how it can be captured by the conceptually simplest approach, a dynamical Lindemann criterion. Such a phenomenological approach can be useful in order to locate a transition semiquantitatively, but it certainly cannot give insight into large-scale properties. The Lindemann criterion will not be used for the total fluctuations of the vortex displacement, but for the relative displacement of neighbouring vortices. In this modified form it has proved to be successful in the static case even for systems which have no true long-range order, see equation (6.51). In order to extract information about the anisotropic nature of the driven state it is instructive to look at the relative displacement fluctuations $`w^2(𝐛)`$ $`=`$ $`\overline{[𝐮(𝐛)𝐮(\mathrm{𝟎})]^2}`$ (7.23) $`=`$ $`w_x^2b_x^2/b^2+w_y^2b_y^2/b^2`$ (7.24) as a function of the orientation of the vector $`𝐛`$ between the ideal position of the neighbours. \[Scheidl and Vinokur 1998a\] (?) evaluated the Lindemann criterion $`w^2(𝐛)c_{\mathrm{Li}}^2a^2`$ for the stability of the lattice and determined the velocity dependence of the melting transition. In this context we will not reproduce the detailed results. But it is interesting to point out that pinning results in contributions $`w_x^2`$ $``$ $`{\displaystyle \frac{1}{v^2}},`$ (7.25a) $`w_y^2`$ $``$ $`{\displaystyle \frac{1}{v}}`$ (7.25b) for the bond fluctuations, which imply that for large drift velocities nearest neighbours separated in $`y`$ direction have larger bond fluctuations than nearest neighbours in $`x`$ direction. Hence this phenomenological criterion indicates that the moving lattice (provided it exists at large drive) will decay into decoupled channels. This conclusion from the Lindemann criterion, which examines displacement fluctuations on short scales, is in agreement with the analysis of fluctuations on large scales described above. Concerning the resulting shift of the dynamic melting temperature it agrees with the finding (7.10) from the incoherent shaking temperature (7.8c). Thus the different approaches, the shaking temperature as measure of the strength of the pinning force, the Lindemann criterion as a measure of the displacement fluctuations on small scales and the renormalization-group results for the displacement fluctuations on large scales give in combination a consistent picture of the physics of driven vortices. ## 8 Summary In this article we have reviewed the influence of weak pinning by point-like impurities on the vortex-line lattice in type-II superconductors. In particular, we have addressed the question to what extent these superconductors display glass-like properties and by which order parameter or correlation functions these properties can be identified. Hereby it is important to distinguish two concepts of order: (i) order in the position of the vortex lines, which allows for a breaking of the continuous translational symmetry that is reflected by a spatial modulation of the magnetic induction; and (ii) the order of the superconducting condensate wave function, which is related to the breaking of the $`U(1)`$ symmetry and which manifests itself in phase coherence (ODLRO). Pinning destroys the positional long-range order of the VLL in $`d4`$. However, for weak pinning (and for sufficiently low temperature, which is always assumed) quasi-long-range positional order persists in $`2<d4`$. This is true also in superconducting films ($`d=2`$) in a parallel field (where the exponent of the algebraic decay depends only logarithmically on the scale), but not in a perpendicular field, where dislocations induce short-ranged positional order. In $`d>2`$ the vortex lattice behaves as an elastic medium for weak disorder, i.e., it is topologically ordered since there are no free dislocations. Then the vortex system can be called an elastic vortex glass. (A superconducting film in a parallel field can also be considered as an elastic glass since dislocations are excluded for geometrical reasons). In elastic vortex glasses vortices are collectively pinned such that vortex motion can take place only due to thermal activation over arbitrarily high barriers. Such barriers are necessary for a vanishing linear resistivity; otherwise the superconductor in the mixed phase would actually be an Ohmic conductor. In the mixed phase vortex fluctuations are decisive not only for the degree of positional order in the vortex system but also for the phase coherence. In pure systems thermal fluctuation destroy ODLRO in $`d4`$ dimensions (if screening is taken into account), including bulk superconductors and superconducting films in a perpendicular field. Preliminary calculations to lowest order in $`ϵ`$ for $`d=4ϵ`$ dimensions suggest that this conclusion remains true also for systems with weak disorder. A phase-coherent vortex glass, where the condensate wave function is quenched, can exist only in unphysical dimensions $`d>4`$. However, higher order terms in $`ϵ`$ may still change this result. This applies in particular to systems with strong disorder (e.g. gauge-glass models without screening) that substantially suppresses thermal fluctuations and may permit the persistence of phase coherence in $`d4`$. When the elastic vortex glass breaks down due to the proliferation of dislocations, it may still possess a weaker degree of order that might disappear only at higher temperatures and/or stronger disorder. Since the positional order of the elastic vortex glass in $`2<d4`$ resembles that of the pure crystal in $`d=2`$, it is possible that the elastic vortex glass breaks into a hexatic vortex liquid with bond-orientational order \[Chudnovsky 1989, Toner 1991b\] before it decays into an isotropic vortex liquid. A more detailed investigation of this possibility as well as of further more exotic disordered phases is left for future studies. ## Acknowledgements During the last decade we had enjoyable and valuable discussions with numerous colleagues. We wish to express our thanks to all of them. We are particularly grateful to S. Bogner, G. Eilenberger, T. Emig, T. Giamarchi, D. A. Huse, J. Kierfeld, M. Lässig, P. Le Doussal, R. Ikeda, M. A. Moore, L. Radzihovsky, H. Rieger, and V. M. Vinokur. In view of the width of the subject and the desired limitation of length of this article it was impossible to include all contributions to the field. We apologize to the authors of these contribution that were paid less tribute than they deserve. ## Appendix A Pinning of periodic media In this appendix we present the transformation of the pinning energy for vortex line lattices, which leads to an effective periodic pinning potential. In our notation $`𝐫=(𝐱,𝐳)`$ denotes a point of space, with the $`D`$-dimensional component $`𝐳`$ component along the vortex lines (in general dimension, a manifold) and the $`N`$-dimensional orthogonal component $`𝐱`$ of the displacement. The dimension of the embedding space of the VLL then is $`d=N+D`$. In the undistorted lattice, the vortex lines are located at positions $`𝐗`$. The distorted lattice can be described by the density (lines per $`N`$-dimensional volume) $$\rho _𝐮(𝐫)=\underset{𝐗^{}}{}\delta (𝐱𝐗^{}𝐮(𝐗^{},𝐳)).$$ (A.1) The average density is denoted by $`\rho _0`$ and it is related to the average vortex spacing $`a`$ through $`\rho _0=a^N`$. In particular, $`\rho _0=B/\mathrm{\Phi }_0`$ for dimension $`N=2`$. The pinning energy reads $`_{\mathrm{pin}}`$ $`=`$ $`{\displaystyle d^Dz\underset{𝐗}{}V(𝐗+𝐮(𝐗,𝐳),𝐳)}={\displaystyle d^dr\rho _𝐮(𝐫)V(𝐫)}`$ (A.2) We always assume disorder to be Gaussian distributed with correlations $$\overline{V(𝐱,𝐳)V(𝐱^{},𝐳^{})}=\mathrm{\Delta }(𝐱𝐱^{})\delta (𝐳𝐳^{}).$$ (A.3) Using the Poisson-summation formula \[Nattermann et al. 1991, Giamarchi and Le Doussal 1995\], we may rewrite the density $`\rho _𝐮(𝐫)`$ $`=`$ $`{\displaystyle \underset{𝐗^{}}{}}{\displaystyle d^Nx^{}\delta (𝐱𝐱^{}𝐮(𝐱^{},𝐳))\delta (𝐗^{}𝐱^{})}`$ (A.4) $`=`$ $`{\displaystyle d^Nx^{}\delta (𝐱𝐱^{}𝐮(𝐱^{},𝐳))\rho _0\underset{𝐐}{}e^{i𝐐𝐱^{}}}`$ $`=`$ $`\mathrm{det}^1[\delta _{\alpha \beta }+_\alpha u_\beta (𝐱,𝐳)]\rho _0{\displaystyle \underset{𝐐}{}}e^{i𝐐[𝐱\stackrel{~}{𝐮}(𝐫)]}`$ $``$ $`[1_\alpha u_\alpha (𝐱,𝐳)]\rho _0{\displaystyle \underset{𝐐}{}}e^{i𝐐[𝐱\stackrel{~}{𝐮}(𝐫)]}`$ $``$ $`\rho _0_\alpha u_\alpha (𝐱,𝐳)+\rho _0{\displaystyle \underset{𝐐}{}}\mathrm{cos}\{𝐐[𝐱\stackrel{~}{𝐮}(𝐫)]\},`$ where $`𝐐`$ are reciprocal lattice vectors (RLV) and $`_\alpha =/x_\alpha `$. The displacement tells us how vortices are shifted form the “ideal” position $`𝐱^{}`$ to the “actual” position $`𝐱`$. In the beginning the displacement is considered as a function of the “ideal” position, $`𝐱𝐱^{}=𝐮(𝐱^{},𝐳)`$. During these manipulation we have expressed the displacement as a function of the actual position, $`𝐱𝐱^{}=\stackrel{~}{𝐮}(𝐱,𝐳)`$. The random potential $`V`$ couples to the divergence of the displacement in the density, Eq. (A.4), as an effective random-compression force. This term is of particular importance in dimensions $`d2`$. The remaining contributions $`𝐐\mathrm{𝟎}`$ to Eq. (A.4) represent an effective periodic pinning potential since it is invariant under shifts $`\stackrel{~}{𝐮}(𝐱,𝐳)\stackrel{~}{𝐮}(𝐱,𝐳)+𝐗`$ with an arbitrary lattice vector $`𝐗`$. However, since under this shift $`𝐱`$ is to be held fixed, the shift means $`𝐱^{}𝐱^{}𝐗`$. Thus it does not mean a translation of the vortex lattice in the laboratory frame but it means as relabeling of the vortices, or a shift of the ideal reference positions $`𝐱^{}`$. In practice it is more convenient to start with the exact pinning energy for a VLL in the replicated system: $`_{\mathrm{pin},n}`$ $`=`$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{a,b}{}}{\displaystyle d^Dz\underset{𝐗,𝐗^{}}{}\mathrm{\Delta }(𝐗+𝐮^a(𝐗,𝐳)𝐗^{}𝐮^b(𝐗^{},𝐳))}`$ (A.5) $`=`$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{a,b}{}}{\displaystyle d^Dz\underset{𝐗,𝐗^{}}{}_𝐤\widehat{\mathrm{\Delta }}(𝐤)e^{i𝐤[𝐗+𝐮^a(𝐗,𝐳)𝐗^{}𝐮^b(𝐗^{},𝐳)]}},`$ where $`_𝐤=d^dk/(2\pi )^d`$ and the Fourier transform of the correlator is $`\widehat{\mathrm{\Delta }}(𝐤){\displaystyle d^Nxe^{i𝐤𝐱}\mathrm{\Delta }(𝐱)}.`$ (A.6) The energy (A.5) can then be transformed in the following way: the wave vector $`𝐤=𝐐+𝐪`$ is split into a reciprocal lattice vector $`𝐐`$ and a vector $`𝐪`$ within the first Brillouin zone and $`_𝐤=_𝐐_𝐪`$, accordingly. The contribution $`𝐐=\mathrm{𝟎}`$ represents the pinning energy for a continuous elastic medium. This contribution contains a random compression force (the terms of order $`q^2`$ in this contribution) which we will not discuss further here. We focus on the contributions $`𝐐\mathrm{𝟎}`$ which encode the periodicity of the VLL. For a short disorder correlation length $`\xi a`$ one may approximate $`\widehat{\mathrm{\Delta }}(𝐐+𝐪)\widehat{\mathrm{\Delta }}(𝐐)`$ and then perform the the integration over $`𝐪`$, after which only the contributions $`𝐗=𝐗^{}`$ survive since $`|𝐮^a(𝐗,𝐳)𝐮^b(𝐗^{},𝐳)||𝐗𝐗^{}|`$ for a roughness exponent $`\zeta <1`$. After these approximations one ends up with $`_{\mathrm{pin},n}`$ $``$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{a,b}{}}{\displaystyle d^Dz\underset{𝐗}{}\rho _0\underset{𝐐}{}\mathrm{\Delta }(𝐐)e^{i𝐐[𝐮^a(𝐗,𝐳)𝐮^b(𝐗,𝐳)]}}`$ (A.7) $`=`$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{a,b}{}}{\displaystyle d^Dz\underset{𝐗}{}a^N\stackrel{~}{\mathrm{\Delta }}(𝐮^a(𝐗,𝐳)𝐮^b(𝐗,𝐳))}`$ $``$ $`{\displaystyle \frac{1}{2T}}{\displaystyle \underset{a,b}{}}{\displaystyle d^dr\stackrel{~}{\mathrm{\Delta }}(𝐮^a(𝐫)𝐮^b(𝐫))}`$ with the effective periodic correlator $`\stackrel{~}{\mathrm{\Delta }}(𝐮)`$ $`=`$ $`\rho _0^2{\displaystyle \underset{𝐐}{}}\mathrm{\Delta }(𝐐)e^{i𝐐𝐮}=\rho _0{\displaystyle \underset{𝐗}{}}\mathrm{\Delta }(𝐗+𝐮),`$ (A.8) which is related to the effective potential via $`\overline{\stackrel{~}{V}(𝐮,𝐫)\stackrel{~}{V}(𝐮^{},𝐫^{})}\stackrel{~}{\mathrm{\Delta }}(𝐮𝐮^{})\delta (𝐫𝐫^{}).`$ (A.9) This means that the periodic lattice effectively behaves like an elastic manifold of internal dimension $`D_{\mathrm{eff}}=D+N`$ in a effective embedding space $`d_{\mathrm{eff}}=D+2N`$ subject to a periodic pinning potential. The advantage of the manipulations (A.7) as compared to (A.4) is that we end up with a periodic correlator for the displacement $`𝐮`$ as a function of the ideal reference position (which is the actual degree of freedom of the vortex lattice) rather than for $`\stackrel{~}{𝐮}`$ as a function of the actual vortex position. ## Appendix B List of recurrent symbols Note that functions and their Fourier transforms have the same name although they are different functions. This ambiguity is removed by the argument of the function, which is either a length (such as $`𝐱`$, $`𝐳`$, $`𝐫`$ etc.) or a wave vector (such as $`𝐤`$, $`𝐪`$, $`𝐐`$ etc.). The only exceptions are $`\mathrm{\Delta }`$ and $`\stackrel{~}{\mathrm{\Delta }}`$. Their common Fourier transform is denoted by $`\widehat{\mathrm{\Delta }}`$.
warning/0003/hep-th0003016.html
ar5iv
text
# Gravitational Waves in Open de Sitter Space ## I Introduction One appeal of inflationary cosmology is its mechanism for the origin of cosmological perturbations. The de Sitter phase of exponentially-rapid expansion quickly redshifts away any local perturbations, leaving behind only the quantummechanical vacuum fluctuations in the various fields. During inflation, these perturbations are stretched to macroscopic length scales and subsequently amplified, to later seed the growth of the large scale structures in the present-day universe. A particularly clean example of this effect are the gravitational wave perturbations of the spacetime itself. These tensor perturbations contribute to the cosmic microwave background anisotropy via the Sachs-Wolfe effect. They may potentially provide an observational discriminant between different theories of open (or closed) inflation because their long-wavelength modes strongly depend on the boundary conditions at the instanton that describes the beginning of the inflationary universe . Although the tensor spectrum has been successfully computed in realistic $`O(3,1)`$ invariant models for an open inflationary universe , the problem of calculating the primordial gravitational waves in perfect open de Sitter spacetime has remained a paradox for some time. The previous literature claims that the spectrum of gravitational waves in perfect de Sitter space is infrared divergent for all physically well-motivated initial quantum states of an eternally inflating universe . Breaking the $`O(4,1)`$ invariance of de Sitter space by going to a realistic inflationary model introduces a potential barrier for the tensor fluctuation modes, and it has been argued that the bubble wall acts to regularise the divergent spectrum in perfect de Sitter space . Previous calculations of the gravitational wave spectrum in open de Sitter space are based on a mode-by-mode analysis. One has a prescription for the vacuum state of the graviton that is imposed on every mode separately, on some Cauchy surface for the de Sitter spacetime. Then one propagates each mode into the open universe region. In this paper we instead compute the two-point tensor correlator in real space. In doing so, we have obtained an infrared finite tensor spectrum. The difference in the two approaches is related to the non-uniqueness of the mode decomposition in an open universe, as we shall explain. As an aside, we mention in this context that also fluctuations of a massless minimally coupled scalar field in de Sitter space do not break $`O(4,1)`$. In some prior literature (see e.g. ) it is shown that there is no de Sitter invariant propagator for such a scalar field. However, the scalar field is not itself an observable since the action depends only on its derivative, and there is a symmetry $`\varphi \varphi +`$ constant. In fact, correlators of space or time derivatives of $`\varphi `$ are de Sitter invariant, and since these are the only physical correlators in the theory, de Sitter invariance is unbroken. We implement the Hartle–Hawking no boundary proposal in our work by ’rounding off’ open de Sitter space on a compact Euclidean instanton, namely a round four sphere. The fluctuations are computed in the Euclidean region directly from the Euclidean path integral, to first order in $`\overline{h}`$ around the instanton saddle point. The Euclidean two-point correlator is analytically continued into the Lorentzian region where it describes the quantum mechanical vacuum fluctuations of the graviton field in the state described by the no boundary proposal initial conditions. There is no ambiguity in the choice of initial conditions because the Euclidean correlator is unique. ## II Tensor Fluctuations about Cosmological Instantons In quantum cosmology the basic object is the wavefunctional $`\mathrm{\Psi }[h_{ij},\varphi ]`$, the amplitude for a three-geometry with metric $`h_{ij}`$ and field configuration $`\varphi `$. It is formally given by a path integral $`\mathrm{\Psi }[h_{ij},\varphi ]{\displaystyle ^{h_{ij},\varphi }}\left[𝒟g\right]\left[𝒟\varphi \right]e^{iS[g,\varphi ]}.`$ (1) Following Hartle and Hawking the lower limit of the path integral is defined by continuing to Euclidean time and integrating over all compact Riemannian metrics $`g`$ and field configurations $`\varphi `$. If one can find a saddle point of (1), namely a classical solution satisfying the Euclidean no boundary condition, one can in principle at least compute the path integral as a perturbative expansion to any desired power in $`\mathrm{}`$. In this paper we wish to compute the two-point tensor fluctuation correlator in open de Sitter spacetime, $$ds^2=dt^2+\mathrm{sinh}^2(t)(d\chi ^2+\mathrm{sinh}^2(\chi )d\mathrm{\Omega }_2^2)).$$ (2) Open de Sitter space may be obtained by analytic continuation of an O(5) invariant instanton, describing the beginning of a semi-eternally inflating universe. The analytic continuation is given by setting $`t=i\sigma `$ and the radial coordinate $`\chi =i\mathrm{\Omega }`$, where $`\mathrm{\Omega }`$ is the polar angle on the three sphere (see ). The instanton obtained in this way is a solution of the Euclidean equations of motion with the maximal symmetry allowed in four dimensions. It takes the form of a round four sphere with line element $`ds^2=d\sigma ^2+\mathrm{sin}^2(\sigma )d\mathrm{\Omega }_3^2`$, where $`d\mathrm{\Omega }_3^2`$ is the line element on $`S^3`$. It is useful to introduce a conformal spatial coordinate $`X`$ defined by $`_\sigma ^{\pi /2}\frac{d\sigma ^{}}{\mathrm{sin}\sigma ^{}}`$, so that the line element takes the form $`ds^2`$ $`=`$ $`\mathrm{cosh}^2X\left(dX^2+d\mathrm{\Omega }_3^2\right).`$ (3) On the four sphere $`X`$ then ranges from $`\mathrm{}`$ to $`+\mathrm{}`$. The principles of our method to calculate cosmological perturbations are described in detail in . The instanton solution provides the classical background with respect to which the quantum fluctuations are defined. In the Euclidean region the exponent $`iS`$ in the path integral becomes $`S_E=(S_0+S_2)`$, where $`S_E`$ is the Euclidean action, $`S_0`$ is the instanton action and $`S_2`$ the action for fluctuations. We keep the latter only to second order. The path integral for the two-point tensor fluctuation about a particular instanton background is then given by $`t_{ij}(x)t_{i^{}j^{}}(x^{})={\displaystyle \frac{\left[𝒟\delta g\right]\left[𝒟\delta \varphi \right]e^{S_2}t_{ij}(x)t_{i^{}j^{}}(x^{})}{\left[𝒟\delta g\right]\left[𝒟\delta \varphi \right]e^{S_2}}}.`$ (4) To first order in $`\overline{h}`$ the quantum fluctuations are specified by a Gaussian integral. The Euclidean action determines the allowed perturbation modes because divergent modes are suppressed in the path integral. The Euclidean two-point tensor correlator is then analytically continued into the Lorentzian region where it describes the quantum mechanical vacuum fluctuations of the graviton field in the state described by the no boundary proposal initial conditions. To find the perturbed action $`S_2`$ that enters in the path integral (4), we write the perturbed line element in open de Sitter space as $$ds^2=\mathrm{sinh}^2(\tau )\left((1+2A)d\tau ^2+S_idx^id\tau +(\gamma _{ij}+h_{ij})dx^idx^j\right),$$ (5) where the fields $`A`$, $`S_i`$ and $`h_{ij}`$ are small perturbations. Because we are interested in the gravitational wave spectrum in the open slicing of de Sitter space, we will only retain $`O(3,1)`$ invariance in our calculation. The quantities $`S_i`$ and $`h_{ij}`$ may be uniquely decomposed as follows , $`h_{ij}`$ $`=`$ $`{\displaystyle \frac{1}{3}}h\gamma _{ij}+2\left(_i_j{\displaystyle \frac{\gamma _{ij}}{3}}\mathrm{\Delta }_3\right)E+2F_{(i|j)}+t_{ij},`$ (6) $`S_i`$ $`=`$ $`B_{|i}+V_i.`$ (7) Here $`\mathrm{\Delta }_3`$ is the Laplacian on $`S^3`$ and $`|j`$ the covariant derivative on the three-sphere. With respect to reparametrisations of the three-sphere, $`h`$, $`B`$ and $`E`$ are scalars, $`V_i`$ and $`F_i`$ are divergenceless vectors and $`t_{ij}`$ is a transverse traceless symmetric tensor, describing the gravitational waves. Because gauge transformations are scalar or vector, the perturbations $`t_{ij}`$ are automatically gauge invariant. It is important to note that the gauge invariance of $`t_{ij}`$ follows from the uniqueness of the above decomposition. This is only true however for bounded (asymptotically decaying) perturbations . If one does not impose suitable asymptotic conditions on the fields, a degeneracy appears between scalar and tensor perturbations that introduces a discrete gauge mode in the tensor spectrum, which plays a crucial role in the divergent behaviour of the correlator. We come back to this point in Section V. We now substitute the decomposition (6) into the Lorentzian action for gravity plus a cosmological constant, $$S=\frac{1}{2\kappa }d^4x\sqrt{g}\left(R2\mathrm{\Lambda }\right)\frac{1}{\kappa }d^3x\sqrt{\gamma }K,$$ (8) The scalar, vector and tensor quantities decouple. Keeping all terms to second order, we continue the perturbed Lorentzian action to the Euclidean region. The scalar and vector fluctuations are pure gauge in perfect de Sitter space. The tensor perturbations $`t_{ij}`$ yield the following well-known positive Euclidean action : $$S_2=\frac{1}{8\kappa }d^4x\frac{\sqrt{\gamma }}{\mathrm{cosh}^2X}\left(t^{ij}t_{ij}^{}+t^{ij|k}t_{ij|k}+2t^{ij}t_{ij}\right).$$ (9) Here prime denotes differentiation with respect to the conformal coordinate $`X`$. After performing the rescaling $`\stackrel{~}{t}_{ij}=\frac{t_{ij}}{\mathrm{cosh}X}`$ and integrating by parts we obtain $$S_2=\frac{1}{8\kappa }d^4x\sqrt{\gamma }\stackrel{~}{t}_{ij}\left(\widehat{K}+3\mathrm{\Delta }_3\right)\stackrel{~}{t}^{ij}+\frac{1}{8\kappa }\left[d^3x\sqrt{\gamma }\stackrel{~}{t}_{ij}\stackrel{~}{t}^{ij}\mathrm{tanh}(X)\right],$$ (10) where the Schrödinger operator $$\widehat{K}=\frac{d^2}{dX^2}\frac{2}{\mathrm{cosh}^2(X)}\frac{d^2}{dX^2}+U(X).$$ (11) Because the fluctuations are specified by a Gaussian integral, we can solve the path integral (4) by looking for the Green function of the operator in its exponent. The potential $`U(X)`$ for the fluctuation modes is well known to be perfectly reflectionless. However, changing its shape slightly would introduce some reflection which becomes increasingly significant at small momenta. Such a change corresponds to breaking the $`O(5)`$ invariance of Euclidean de Sitter space and is exactly what happens in the O(4) invariant Hawking–Turok and Coleman–De Luccia instantons that describe the beginning of realistic open inflationary universes. This difference between both classes of instantons has profound implications for the tensor perturbations about them, especially for their long-wavelength regime . The operator $`\widehat{K}`$ has in all three cases a positive continuum starting at eigenvalue $`p^2=0`$, as well as a single bound state $`\stackrel{~}{t}_{ij}=b(X)q_{ij}(\mathrm{\Omega })`$ at $`p=i`$ which turns out to be a trivial gauge mode. ## III The Euclidean Green Function To evaluate the path integral (4), we first look for the Green function $`G_E^{iji^{}j^{}}(X,X^{},\mathrm{\Omega },\mathrm{\Omega }^{})`$ of the operator in (10). The Euclidean fluctuation correlator (4) will then be given by $`\mathrm{cosh}(X)\mathrm{cosh}(X^{})G_E^{iji^{}j^{}}`$. The Euclidean Green function satisfies $$\frac{1}{4\kappa }\left(\widehat{K}+3\mathrm{\Delta }_3\right)G_{Ei^{}j^{}}^{ij}(X,X^{},\mathrm{\Omega },\mathrm{\Omega }^{})=\delta (XX^{})\gamma ^{\frac{1}{2}}\delta _{i^{}j^{}}^{ij}(\mathrm{\Omega }\mathrm{\Omega }^{}).$$ (12) If we think of the scalar product as defined by integration over $`S^3`$ and summation over tensor indices, then the right hand side is the normalised projection operator onto transverse traceless tensors on $`S^3`$. The Green function $`G_{Ei^{}j^{}}^{ij}`$ can only be a function of the geodesic distance $`\mu (\mathrm{\Omega },\mathrm{\Omega }^{})`$ if it is to be invariant under isometries of the three-sphere. This suggests that $$G_{Ei^{}j^{}}^{ij}(\mu ,X,X^{})=4\kappa \underset{p=3i}{\overset{+i\mathrm{}}{}}G_p(X,X^{})W_{(p)i^{}j^{}}^{ij}(\mu ),$$ (13) where $`W_{(p)i^{}j^{}}^{ij}(\mu )`$ is a bitensor that is invariant under the isometry group O(4). It equals the sum (A3) of the normalised rank-two tensor eigenmodes with eigenvalue $`\lambda _p=p^2+3`$ of the Laplacian on $`S^3`$. Note that the indices $`i,j`$ lie in the tangent space over the point $`\mathrm{\Omega }`$ while the indices $`i^{},j^{}`$ lie in the tangent space over the point $`\mathrm{\Omega }^{}`$. On $`S^3`$ we have $$\mathrm{\Delta }_3W_{(p)i^{}j^{}}^{ij}(\mu )=\lambda _pW_{(p)i^{}j^{}}^{ij}(\mu ).$$ (14) The motivation for the unusual labelling of the eigenvalues of the Laplacian is that, as demonstrated in the Appendix, in terms of the label $`p`$ the bitensor on $`S^3`$ has precisely the same formal expression as the corresponding bitensor on $`H^3`$. It is precisely this property that will enable us in Section IV to continue the Green function from the Euclidean instanton into open de Sitter space without decomposing it in Fourier modes. The relation between the bitensors on $`S^3`$ and $`H^3`$ together with some useful formulae and properties of maximally symmetric bitensors are given in Appendix A. Since the tensor eigenmodes of the Laplacian on $`S^3`$ form a complete basis, we can also write $$\gamma ^{\frac{1}{2}}\delta _{i^{}j^{}}^{ij}(\mathrm{\Omega }\mathrm{\Omega }^{})=\underset{p=3i}{\overset{+i\mathrm{}}{}}W_{(p)i^{}j^{}}^{ij}(\mu (\mathrm{\Omega },\mathrm{\Omega }^{})).$$ (15) Hence by substituting our ansatz (13) for the Green function into (12) we obtain an equation for the X-dependent part of the Green function, $$\left(\widehat{K}p^2\right)G_p(X,X^{})=\delta (XX^{}).$$ (16) The solution to equation (16) is $`G_p(X,X^{})`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Delta }_p}}\left[\mathrm{\Psi }_p^r(X)\mathrm{\Psi }_p^l(X^{})\mathrm{\Theta }(XX^{})+\mathrm{\Psi }_p^l(X)\mathrm{\Psi }_p^r(X^{})\mathrm{\Theta }(X^{}X)\right].`$ (17) $`\mathrm{\Psi }_p^l(X)`$ is the solution to the Schrödinger equation that tends to $`e^{ipX}`$ as $`X\mathrm{}`$, and $`\mathrm{\Psi }_p^r(X)`$ is the solution going as $`e^{ipX}`$ as $`X+\mathrm{}`$. The factor $`\mathrm{\Delta }_p`$ is the Wronskian of the two solutions. Since the potential is reflectionless on the round four sphere the left- and right-moving waves do not mix and they equal the Jost functions $`g_{\pm p}(X)`$ with nice analytic properties. The solutions may be found explicitely and are given by $$\{\begin{array}{ccc}\mathrm{\Psi }_p^r(X)\hfill & =\hfill & (\mathrm{tanh}Xip)e^{ipX}\hfill \\ \mathrm{\Psi }_p^l(X)\hfill & =\hfill & (\mathrm{tanh}X+ip)e^{ipX}\hfill \end{array}$$ (18) and their Wronskian $`\mathrm{\Delta }_p=2ip(1+p^2)`$, independent of $`X`$. The zero of the Wronskian at $`p=i`$ corresponds to the bound state mentioned above. Taking $`X>X^{}`$, we obtain the Euclidean Green function as a discrete sum $$G_E^{iji^{}j^{}}(\mu ,X,X^{})=4\kappa \underset{p=3i}{\overset{i\mathrm{}}{}}\frac{i}{2p}\frac{\mathrm{\Psi }_p^r(X)\mathrm{\Psi }_p^l(X^{})}{(1+p^2)}W_{(p)}^{iji^{}j^{}}(\mu ).$$ (19) Before proceeding, let us demonstrate that the Euclidean Green function is regular at the poles of the four sphere. This is a nontrivial check because the coordinates $`\sigma `$ and $`X`$ are singular there, and the rescaling becomes divergent too. In the large $`X,X^{}`$ limit, (19) becomes $$G_E^{iji^{}j^{}}(\mu ,X,X^{})=2\kappa \underset{n=3}{\overset{\mathrm{}}{}}\frac{1}{n}e^{n(XX^{})}W_{(in)}^{iji^{}j^{}}(\mu )$$ (20) For $`n3`$ the Gaussian hypergeometric functions $`F(3+n,3n,7/2,z)`$ that constitute the bitensor $`W_{(n)}^{iji^{}j^{}}`$ have a series expansion that terminates, and they essentially reduce to Gegenbauer polynomials $`C_{n3}^{(3)}(12z)`$. Using then the identity $$\underset{l=0}{\overset{\mathrm{}}{}}C_l^\nu (x)q^l=\left(12xq+q^2\right)^\nu $$ (21) with $`q=e^{(XX^{})}`$, one easily sees that the sum (20) indeed converges. We have the Euclidean Green function defined as an infinite sum (19). However, the eigenspace of the Laplacian on $`H^3`$ suggests that the Lorentzian Green function is most naturally expressed as an integral over real $`p`$. To do so we must extend the summand into the upper half $`p`$-plane. We have already defined the wavefunctions $`\mathrm{\Psi }_p(X)`$ as analytic functions for all complex $`p`$ but we need to extend the bitensor as well. When the Green function is expressed as a discrete sum, it involves the bitensor $`W_{(p)}^{iji^{}j^{}}(\mu )`$ evaluated at $`p=ni`$ with $`n`$ integral. At these values of $`p`$, the bitensor is regular at both coincident and opposite points on $`S^3`$, that is at $`\mu =0`$ and $`\mu =\pi `$. However, if we extend $`p`$ into the complex plane we lose regularity at $`\mu =0`$, essentially because the bitensor obeys the differential equation (12) with a delta function source at $`\mu =0`$. Similarly we must maintain regularity at $`\mu =\pi `$, since there is no delta function source there. The condition of regularity at $`\pi `$ imposed by the differential equation for the Green function is sufficient to uniquely specify the analytic continuation of $`W_{(in)}^{iji^{}j^{}}(\mu )`$ into the complex $`p`$-plane. The continuation is described in the Appendix, and the extended bitensor $`W_{(p)}^{iji^{}j^{}}(\mu )`$ is defined by equations (A6) and (A12). Now we are able to write the sum in (19) as an integral along a contour $`𝒞_1`$ encircling the points $`p=3i,4i,\mathrm{}Ni`$, where $`N`$ tends to infinity. For $`X>X^{}`$ we have $`G_E^{iji^{}j^{}}(\mu ,X,X^{})`$ $`=`$ $`\kappa {\displaystyle _{𝒞_1}}{\displaystyle \frac{dp}{p\mathrm{sinh}p\pi }}{\displaystyle \frac{\mathrm{\Psi }_p^r(X)\mathrm{\Psi }_p^l(X^{})}{(1+p^2)}}W_{(p)}^{iji^{}j^{}}(\mu ).`$ (22) To see that (22) is equivalent to the sum (19) introduce $`1=\mathrm{cosh}p\pi /\mathrm{cosh}p\pi `$ into the integral. Then note that $`\mathrm{coth}p\pi `$ has residue $`\pi ^1`$ at every integer multiple of $`i`$. Finally, use (A15) to rewrite $`W_{(p)}^{iji^{}j^{}}(\mu )`$ in the form regular at $`\mu =0`$ used in (19). The factor of $`\mathrm{cosh}p\pi `$ from (A15) cancels that in the integrand. We now distort the contour for the $`p`$ integral to run along the real $`p`$ axis (Figure 1). At large imaginary $`p`$ the integrand decays exponentially and the contribution vanishes in the limit of large $`N`$. However as we deform the contour towards the real axis we encounter two poles in the $`\mathrm{sinh}p\pi `$ factor, the latter at $`p=i`$ becoming a double pole due to the simple zero of the Wronskian. For the $`p=2i`$ pole, it follows from the normalisation of the tensor harmonics that $`W_{(2i)}^{iji^{}j^{}}=0`$. Indirectly, this is a consequence of the fact that spin-2 perturbations do not have a monopole or dipole component. At $`p=i`$ we have a double pole, but although the relevant Schrödinger operator possesses a bound state, it does not generate a ‘super-curvature mode’. Instead the relevant mode is a time-independent shift in the metric perturbation which may be gauged away . We conclude that up to a term involving a pure gauge mode, we can deform the contour $`𝒞_1`$ into the contour shown in Figure 1. For the moment, since the integrand involves a factor $`p\mathrm{sinh}p\pi `$ which has a double pole at $`p=0`$, we leave the contour avoiding the origin on a small semicircle in the upper half $`p`$-plane. Finally, in order to deal with the pole at $`p=0`$, we re-express the integrand in (22) as a sum of its $`p`$-symmetric and $`p`$-antisymmetric parts. Denoting the integrand by $`I_p`$ we then have $$G_E^{iji^{}j^{}}=\frac{1}{2}𝑑p(I_p+I_p)+\frac{1}{2}𝑑p(I_pI_p),$$ (23) where the integral is taken from $`p=\mathrm{}`$ to $`\mathrm{}`$ along a path avoiding the origin above. But $`𝑑pI_p`$ along this contour is equal to the integral of $`I_p`$ taken along a contour avoiding the origin below. The second term is therefore equal to the integral of $`I_p`$ along a contour around the origin. Hence we have $$\frac{1}{2}𝑑p(I_pI_p)=\pi i\mathrm{𝐑𝐞𝐬}(I_p;p=0).$$ (24) We defer a detailed discussion of this term to Section V, because its interpretation is clearer in the Lorentzian region. Hence for the time being we just keep it, but it will turn out that it represents a non-physical contribution to the graviton propagator. In the $`p`$-symmetric part of the correlator, we can leave the integrand as a sum of $`I_p`$ and $`I_p`$. We henceforth denote the path from $`\mathrm{}`$ to $`+\mathrm{}`$ avoiding the origin above by $``$. This shall turn out to be a regularised version of the integral over the real axis. Our final result for the Euclidean Green function then reads $`G_{iji^{}j^{}}^E(\mu ,X,X^{})`$ $`=`$ $`{\displaystyle \frac{\kappa }{2}}{\displaystyle _{}}{\displaystyle \frac{dp}{p\mathrm{sinh}p\pi }}{\displaystyle \frac{W_{iji^{}j^{}}^{(p)}(\mu )}{(1+p^2)}}\left(\mathrm{\Psi }_p(X)\mathrm{\Psi }_p(X^{})+\mathrm{\Psi }_p(X)\mathrm{\Psi }_p(X^{})\right)`$ (26) $`\pi i\mathrm{𝐑𝐞𝐬}(I_p;p=0).`$ ## IV Two-Point Tensor Correlator in Open de Sitter Space The analytic continuation into open de Sitter space is given by setting $`\sigma =it`$ and the polar angle $`\mathrm{\Omega }=i\chi `$. Without loss of generality we may take one of the two points, say $`\mathrm{\Omega }^{}`$ to be at the north pole of the three-sphere. Then $`\mu =\mathrm{\Omega }`$, and $`\mu `$ continues to $`i\chi `$. We then obtain the correlator in open de Sitter space where one point has been chosen as the origin of the radial coordinate $`\chi `$. The conformal coordinate $`X`$ continues to conformal time $`\tau `$ as $`X=\tau \frac{i\pi }{2}`$ (see ). Hence the analytic continuation of the Euclidean mode functions is given by $`\mathrm{\Psi }_p^r(X)e^{\frac{p\pi }{2}}\mathrm{\Psi }_p^L(\tau )\mathrm{and}\mathrm{\Psi }_p^l(X)e^{\frac{p\pi }{2}}\mathrm{\Psi }_p^L(\tau )`$ (27) where the Lorentzian mode functions are $$\mathrm{\Psi }_p^L(\tau )=(\mathrm{coth}\tau +ip)e^{ip\tau }.$$ (28) They are solutions to the Lorentzian perturbation equation $`\widehat{K}\mathrm{\Psi }_p^L(\tau )=p^2\mathrm{\Psi }_p^L(\tau )`$. In order to perform the substitution $`\mu =i\chi `$, where $`\chi `$ is the comoving separation on $`H^3`$, we use the explicit formula given in the appendix for the bitensor regular at $`\mu =\pi `$. The continued bitensor $`W_{iji^{}j^{}}^{(p)}(\chi )`$ is defined by the equations (A12), (A19) and (A25). It can be seen from (A25) that it involves terms which behave as $`e^{\pm p(i\chi +\pi )}`$. One must extract the $`e^{p\pi }`$-factors in order for the bitensor to correspond to the usual sum of rank-two tensor harmonics on the real $`p`$-axis. To do so we use the following general identity. For $`\tau ^{}\tau >0`$, we have (up to the $`p=i`$ gauge mode) $`{\displaystyle _𝒞}{\displaystyle \frac{dp}{p}}{\displaystyle \frac{\mathrm{\Psi }_p^L(\tau )\mathrm{\Psi }_p^L(\tau ^{})}{(1+p^2)}}e^{ip\chi }F(p)=0,`$ (29) where $`F(p)`$ are the $`p`$-dependent coefficients occurring in the final (Lorentzian) form of the bitensor given in (A33). This identity follows from the analyticity of the integrand. By inserting $`1=\mathrm{sinh}p\pi /\mathrm{sinh}p\pi `$ under the integral, it is clear that the integral (29) with a factor $`e^{p\pi }/\mathrm{sinh}p\pi `$ inserted equals that with a factor $`e^{p\pi }/\mathrm{sinh}p\pi `$ inserted. The resulting identity allows us to replace the factors $`e^{+p(i\chi +\pi )}`$ in the bitensor by $`e^{p(i\chi \pi )}`$, and vice versa in the analog integral of $`I_p`$ closed in the lower half $`p`$-plane. For the tensor correlator we also need to restore the factor $`ia^1(\tau )`$ to $`t_{ij}`$. It is convenient to define the eigenmodes $`\mathrm{\Phi }_p^L(\tau )=\mathrm{\Psi }_p^L(\tau )/a(\tau )`$. The extra minus sign hereby introduced is cancelled by a change in sign of the normalisation factor $`Q_p`$ of the bitensor, which then becomes $`+(p^2+4)/(30\pi ^2)`$. This corresponds to requiring the spacelike metric to have postive signature. We finally obtain the Lorentzian tensor Feynman (time-ordered) correlator, for $`\tau ^{}\tau >0`$, $`t_{ij}(x),t_{i^{}j^{}}(x^{})`$ $`=`$ $`{\displaystyle \frac{\kappa }{2}}{\displaystyle _R}{\displaystyle \frac{dp}{p\mathrm{sinh}p\pi }}{\displaystyle \frac{W_{iji^{}j^{}}^{L(p)}(\chi )}{(1+p^2)}}\left(e^{p\pi }\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})+e^{p\pi }\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})\right)`$ (31) $`\pi i\mathrm{𝐑𝐞𝐬}(I_p^L;p=0),`$ where the Lorentzian bitensor $`W_{iji^{}j^{}}^{L(p)}`$ is defined in the Appendix, equations (A6) and (A33). In this section, we concentrate on the first term in (26), the integral over $`p`$, and ignore for the moment the second, discrete term. We first extract the symmetrised part, $`\{t_{ij}(x),t_{i^{}j^{}}(x^{})\}`$, which is just the real part of the Feynman correlator. The imaginary part involves an integrand which is analytic for $`p0`$: $`t_{ij}(x),t_{i^{}j^{}}(x^{})`$ $`=`$ $`{\displaystyle \frac{\kappa }{2}}{\displaystyle _{}}{\displaystyle \frac{dp}{p(1+p^2)}}W_{iji^{}j^{}}^{L(p)}(\chi )\mathrm{coth}p\pi [\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})+\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})]`$ (33) $`2\kappa {\displaystyle _0^{\mathrm{}}}𝑑p{\displaystyle \frac{W_{iji^{}j^{}}^{L(p)}(\chi )}{(1+p^2)}}\left[{\displaystyle \frac{1}{p}}\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})\right].`$ It is straightforward to see that if we apply the Lorentzian version of the perturbation operator $`\widehat{K}`$ to (33) with an appropriate heaviside function of $`\tau \tau ^{}`$, the imaginary term will produce the Wronskian of $`\mathrm{\Phi }_p^L(\tau )`$ and $`\mathrm{\Phi }_p^L(\tau )`$, which is proportional to $`ip`$, times $`\delta (\tau \tau ^{})`$. Then the integral over $`p`$ produces a spatial delta function. From this one sees that our Feynman correlator obeys the correct second order partial differential equation, with a delta function source. The delta function source term in (12) goes from being real in the Euclidean region to imaginary in the Lorentzian region because the factor $`\sqrt{g}`$ continues to $`i\sqrt{g}`$. The integral in (31) diverges as $`p^2`$ for $`p0`$, in contrast with realistic models for inflationary universes where a reflection term in (33) regularises the spectrum . However, as we immediately show, even in perfect de Sitter space the integral over $`p`$ is perfectly finite. We rewrite the symmetrised correlator as an integral over real $`0p\mathrm{}`$ as follows. Because the integrand in (33) is even in $`p`$, we have $`\{t_{ij}(x),t_{i^{}j^{}}(x^{})\}`$ $`=`$ $`2\kappa {\displaystyle _ϵ^{\mathrm{}}}{\displaystyle \frac{dp}{\pi p^2}}{\displaystyle \frac{p\pi \mathrm{coth}p\pi }{(1+p^2)}}\mathrm{}\left[\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})\right]W_{iji^{}j^{}}^{L(p)}(\chi )`$ (35) $`{\displaystyle \frac{2\kappa }{\pi ϵ}}\mathrm{\Phi }_0^L(\tau )\mathrm{\Phi }_0^L(\tau ^{})W_{iji^{}j^{}}^{L(0)}(\chi )+O(ϵ),`$ the second term being the contribution from the small semicircle around $`p=0`$. Both terms may be combined under one integral. The resulting integrand is *analytic* as $`p0`$ and one can safely take the limit $`ϵ0`$. The symmetrised correlator is then given by $`\{t_{ij}(x),t_{i^{}j^{}}(x^{})\}=`$ $$2\kappa _0^{\mathrm{}}\frac{dp}{\pi p^2}\left(\frac{p\pi \mathrm{coth}p\pi }{(1+p^2)}\mathrm{}\left[\mathrm{\Phi }_p^L(\tau )\mathrm{\Phi }_p^L(\tau ^{})\right]W_{iji^{}j^{}}^{L(p)}(\chi )\mathrm{\Phi }_0^L(\tau )\mathrm{\Phi }_0^L(\tau ^{})W_{iji^{}j^{}}^{L(0)}(\chi )\right),$$ (36) where the Lorentzian bitensor $`W_{iji^{}j^{}}^{L(p)}`$ is defined in the Appendix, equations (A6) and (A33). In this integral it may be written as $$W_{iji^{}j^{}}^{L(p)}(\chi )=\underset{𝒫lm}{}q_{ij}^{(p)𝒫lm}(\mathrm{\Omega })q_{i^{}j^{}}^{(p)𝒫lm}(\mathrm{\Omega }^{})^{}.$$ (37) The functions $`q_{ij}^{(p)𝒫lm}(\mathrm{\Omega })`$ are the rank-two tensor eigenmodes with eigenvalues $`\lambda _p=(p^2+3)`$ of the Laplacian on $`H^3`$. Here $`𝒫=e,o`$ labels the parity, and $`l`$ and $`m`$ are the usual quantum numbers on the two-sphere. At large $`p`$, the coefficient functions $`w_j^{(p)}`$ of the bitensor (see Appendix A) behave like $`p\mathrm{sin}p\chi `$. Hence the above integral converges at large $`p`$, for both timelike and spacelike separations. Furthermore, the correlations asymptotically decay for large separation of the two points. Equation (31), with the first term given by (36) is our final result for the two-point tensor correlator in open de Sitter space, with Euclidean no boundary initial conditions. Contracting the propagator with the harmonics $`q_{(p)elm}^{i^{}j^{}}`$ and integrating over the three sphere reveals that the second term leaves the spectrum completely unchanged apart from cancelling the (divergent) contribution from the $`p^2=0`$ divergence in the first term. We defer a detailed discussion of this result to the next section, in which we will also clarify the difficulties of the previous work on the graviton propagator in open de Sitter spacetime . As an illustration let us compute the Sachs-Wolfe integral and show that all the multipole moments are finite. The contribution of gravitational waves to the CMB anisotropy in perfect de Sitter space is given by $`{\displaystyle \frac{\delta T_{SW}^{}}{T}}(\theta ,\varphi )`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _0^{\tau _0}}𝑑\tau t_{\chi \chi ,\tau }^{}(\tau ,\chi ,\theta ,\varphi )|_{\chi =\tau _0\tau },`$ (38) where $`\tau _0`$ is the observing time. The temperature anisotropy on the sky is characterised by the two-point angular correlation function $`C(\gamma )`$, where $`\gamma `$ is the angle between two points located on the celestial sphere. It is customary to expand the correlation function in terms of Legendre polynomials as $$C(\gamma )=\frac{\delta T}{T}(0)\frac{\delta T}{T}(\gamma )=\underset{l=2}{\overset{\mathrm{}}{}}\frac{2l+1}{4\pi }C_lP_l(\mathrm{cos}\gamma ).$$ (39) Hence, inserting the Sachs-Wolfe integral into (39) and substituting (36) for the two-point fluctuation correlator yields the multipole moments $`C_l`$ $`=`$ $`{\displaystyle \frac{\kappa }{2}}{\displaystyle _0^+\mathrm{}}dp{\displaystyle _0^{\tau _0}}d\tau {\displaystyle _0^{\tau _0}}d\tau ^{}({\displaystyle \frac{\mathrm{coth}p\pi }{p(1+p^2)}}\mathrm{}\left[\dot{\mathrm{\Phi }}_p^L(\tau )\dot{\mathrm{\Phi }}_p^L(\tau ^{})\right]Q_{\chi \chi }^{pl}Q_{\chi ^{}\chi ^{}}^{pl}`$ (41) $`\dot{\mathrm{\Phi }}_0^L(\tau )\dot{\mathrm{\Phi }}_0^L(\tau ^{})Q_{\chi \chi }^{0l}Q_{\chi ^{}\chi ^{}}^{0l}).`$ In this expression we have written the normalised tensor harmonics $`q_{\chi \chi }^{(p)elm}(\chi ,\theta ,\varphi )`$ as $`Q_{\chi \chi }^{pl}(\chi )Y_{lm}(\theta ,\varphi )`$, where $$Q_{\chi \chi }^{pl}(\chi )=\frac{N_l(p)}{p^2(p^2+1)}(\mathrm{sinh}\chi )^{l2}\left(\frac{1}{\mathrm{sinh}\chi }\frac{d}{d\chi }\right)^{l+1}(\mathrm{cos}p\chi )$$ (42) and $$N_l(p)=\left[\frac{(l1)l(l+1)(l+2)}{\pi _{j=2}^l(j^2+p^2)}\right]^{1/2}.$$ (43) It can readily be seen that the multipole moments are finite. With the aid of the explicit expressions and the wavefunctions (28) they can be numerically computed. ## V Conclusions We have computed the spectrum of primordial gravitational waves predicted in open de Sitter space, according to Euclidean no boundary initial conditions. The Euclidean path integral unambiguously specifies the tensor fluctuations with no additional assumptions. The real space Euclidean correlator has been analytically continued into the Lorentzian region without Fourier decomposing it, and we obtained an infrared finite two-point tensor correlator in open de Sitter space, contrary to previous results in the literature . Let us now elaborate on the second, regularising term in the symmetrised correlator (36) and the discrete $`p=0`$ contribution to the Feynman correlator given from the last term in (24). Not surprisingly, they have a similar interpretation. Their angular part $`W_{iji^{}j^{}}^{L(0)}(\chi )`$ is equal to the sum of the tensor harmonics with eigenvalue $`\lambda _p(p=0)=3`$ of the Laplacian on $`H^3`$. It has been known that a degeneracy appears between $`p^2=0`$ tensor modes and $`p_s^2=4`$ scalar harmonics . More specifically, one has $`q_{ij}^{e(0)lm}=\left(_i_j\frac{1}{3}\gamma _{ij}^2\right)q^{(2i)lm}`$ where $`q^{(2i)lm}=P_{(2i)lm}Y_{lm}`$. The discrete $`p^2=0`$ tensor harmonics are the only transverse traceless tensor perturbations that can be constructed from a scalar quantity. But as a consequence of this, they are sensitive to scalar gauge transformations. Consider now the coordinate transformation $`\xi ^\alpha =(0,ϵ\mathrm{\Phi }_0^L(\tau )^iq^{(2i)lm})`$. Under this transformation the transverse traceless part of the metric perturbation $`h_{ij}`$ in the perturbed line element (5) changes exactly by $`ϵt_{ij}^{(0)lm}=ϵ\mathrm{\Phi }_0^L(\tau )q_{ij}^{(0)lm}`$. Using the transverse-traceless properties of $`t_{ij}`$ it is easily seen that the action for tensor fluctuations is invariant under such transformations. Hence this tensor eigenmode is non-physical and can be gauged away. Note that since the functional form of $`\xi `$ is completely fixed this corresponds to a global transformation, analogous to the transformation $`\varphi \varphi +`$ constant for a massless field. To compute the Green function for a massless field one has to project out this homogeneous mode, and it is necessary to do the same here. One should therefore disregard the contribution from the discrete term in (24) to the Lorentzian correlator. This was actually also done in our computation of the tensor fluctuation spectrum about $`O(4)`$ instantons , although in that case not because the mode was pure gauge, but because it couples to the inflaton field, and is not represented by a simple action of the form (8). If a scalar field is present, the mode is most simply treated as a part of the scalar perturbations, as was done in . In our result (36) for the symmetrised correlator, the discrete gauge mode is set to zero because the second term cancels exactly the contribution from the $`p^2=0`$ mode implicitly contained in the continuous spectrum. This automatic cancellation does not happen in the conventional mode-by-mode analysis where, if one chooses the most degenerate continuous representation of the isometry group $`O(3,1)`$ of the hyperboloid $`H^3`$, corresponding to the range $`p[0,\mathrm{})`$, one obtains a divergent correlator. It is clear that the underlying reason for these subtleties has to do with the different nature of tensor harmonics on compact and non-compact spaces. Hence, we could have expected the generation of the two discrete gauge modes simply from the analytic continuation of the completeness relation (14) of the harmonics on $`S^3`$. Apart from the sum of the complete set of modes that constitute the delta function on $`H^3`$, one obtains also three extra terms $`W_{(2i)}^{iji^{}j^{}}(\mu )`$, $`W_{(i)}^{iji^{}j^{}}(\mu )`$ and $`W_{(0)}^{iji^{}j^{}}(\mu )`$. The first term is zero, and the remaining two terms should respectively be viewed as sums of vector - and scalar harmonics. On the other hand, the fact that the scalar/tensor degeneracy appears precisely at the lower bound of the continuous spectrum is a peculiar feature of three dimensions. In the analogous computation in four dimensions for instance , this degeneracy happens at $`p^2=1/4`$ and consequently, there is no regularising term in the correlator. There is yet another way in which the exclusion of the degenerate modes from the perturbation spectrum can be interpreted. Remember that in non-compact spacetimes the decomposition (6) is only uniquely defined for bounded perturbations. Hence, the only way there can appear a degeneracy between the different types of fluctuations is for the degenerate modes to be unbounded. Indeed, on the three-hyperboloid the scalar $`p^2=4`$ modes describe divergent fluctuations because the scalar spherical harmonics $`q^{(2i)lm}`$ grow exponentially with distance. The action of the above tensor operator renders only the $`q_{\chi j}^{(0)lm}`$ components of $`q_{ij}^{(0)lm}`$ finite at infinity. The remaining components still diverge as $`e^\chi `$ and correspond to exponentially growing fluctuations at large distances<sup>*</sup><sup>*</sup>*The confusion arises because, due to the form of the metric inverse, scalar invariants are finite at infinity, e.g. $`q_{ij}q^{ij}e^{2\chi }`$. This also explains why the coefficient functions $`w_j^{(0)}(\chi )`$ in the bitensor $`W_{iji^{}j^{}}^{L(0)}`$ asymptotically decay.. Since in cosmological perturbation theory one assumes the perturbation $`h_{ij}`$ to be small, one must expand correlators in bounded harmonics. We want to emphasize that the regularity of the two-point tensor correlator does not depend on the Euclidean methods used in our work. One could have equally well computed the correlator on closed Cauchy surfaces for the de Sitter space where the subtleties encountered here do not arise, assuming the standard conformal vacuum for that slicing. One would then analytically continue the result to the open slicing. On the other hand, the Euclidean no boundary principle is an appealing prescription which avoids the arbitrary choice of vacuum otherwise needed. The path integral effectively defines its own initial conditions, yielding a unique and infrared finite Green function in the Lorentzian region. The initial quantum state of the perturbation modes, defined by the no boundary path integral, corresponds to the conformal vacuum in the Lorentzian spacetime. This is in many ways the most natural state in de Sitter space, but the regularity of the graviton propagator is independent of this choice. The most important technical advantage of our method is that we deal throughout directly with the real space correlator, which makes the derivation independent of the gauge ambiguities involved in the mode decomposition. Finally, let us conclude by comparing the gravitational wave spectrum in perfect open de Sitter spacetime with the spectrum in realistic open inflationary universes. In both the Hawking–Turok and the Coleman–De Luccia model for open inflation there is an extra reflection term in the correlator because O(5) symmetry is broken on the instanton . This term gives rise to long-wavelength bubble wall fluctuations in the Lorentzian region. At first sight, the wall fluctuations seem to regularise the spectrum. However, adding and subtracting the second term in (36) to the two-point tensor correlator in the $`O(4)`$ models (eq. (34) in ) and comparing that with our result (36) reveals that the wall fluctuations actually appear as an extra long-wavelength continuum contribution *on top of* the spectrum in perfect de Sitter space. Hence in both the Hawking–Turok and Coleman–De Luccia model there is an enhancement of the fluctuations compared to the perturbations in perfect de Sitter space. But the singularity in Hawking–Turok instantons suppresses the wall fluctuations because it enforces Dirichlet boundary conditions on the perturbation modes . Hence we expect the spectrum in perfect de Sitter space to be quite similar to the spectrum predicted by singular instantons. On the other hand, Coleman–De Luccia models typically predict large wall fluctuations, yielding a very different CMB anisotropy spectrum on large angular scales. The tensor fluctuation spectrum therefore potentially provides an observaional discriminant between different theories of open inflation . Acknowledgements It is a pleasure to thank Steven Gratton and Valery Rubakov for stimulating discussions. ## A Maximally Symmetric Bitensors A maximally symmetric bitensor $`T`$ is one for which $`\sigma ^{}T=0`$ for any isometry $`\sigma `$ of the maximally symmetric manifold. Any maximally symmetric bitensor may be expanded in terms of a complete set of ’fundamental’ maximally symmetric bitensors with the correct index symmetries. For instance $`T_{iji^{}j^{}}`$ $`=`$ $`t_1(\mu )g_{ij}^{}g_{i^{}j^{}}^{}+t_2(\mu )\left[n_i^{}g_{ji^{}}^{}n_j^{}^{}+n_j^{}g_{ii^{}}^{}n_j^{}^{}+n_i^{}g_{jj^{}}^{}n_i^{}^{}+n_j^{}g_{ij^{}}^{}n_i^{}^{}\right]`$ (A2) $`+t_3(\mu )\left[g_{ii^{}}^{}g_{jj^{}}^{}+g_{ji^{}}^{}g_{ij^{}}^{}\right]+t_4(\mu )n_i^{}n_j^{}n_i^{}^{}n_j^{}^{}+t_5(\mu )\left[g_{ij}^{}n_i^{}^{}n_j^{}^{}+n_i^{}n_j^{}g_{i^{}j^{}}^{}\right]`$ where the coefficient functions $`t_j(\mu )`$ depend only on the distance $`\mu (\mathrm{\Omega },\mathrm{\Omega }^{})`$ along the shortest geodesic from $`\mathrm{\Omega }`$ to $`\mathrm{\Omega }^{}`$. $`n_i^{}^{}(\mathrm{\Omega },\mathrm{\Omega }^{})`$ and $`n_i^{}(\mathrm{\Omega },\mathrm{\Omega }^{})`$ are unit tangent vectors to the geodesics joining $`\mathrm{\Omega }`$ and $`\mathrm{\Omega }^{}`$ and $`g_{ij^{}}(\mathrm{\Omega },\mathrm{\Omega }^{})`$ is the parallel propagator along the geodesic; $`V^ig_i^j^{}`$ is the vector at $`\mathrm{\Omega }^{}`$ obtained by parallel transport of $`V^i`$ along the geodesic from $`\mathrm{\Omega }`$ to $`\mathrm{\Omega }^{}`$ . The set of tensor eigenmodes on $`S^3`$ or $`H^3`$ forms a representation of the symmetry group of the manifold. It follows in particular that their sum over the parity states $`𝒫=\{e,o\}`$ and the quantum numbers $`l`$ and $`m`$ on the two-sphere defines a maximally symmetric bitensor on $`S^3`$ (or $`H^3`$) $$W_{(p)i^{}j^{}}^{ij}(\mu )=\underset{𝒫lm}{}q_{𝒫lm}^{(p)ij}(\mathrm{\Omega })q_{i^{}j^{}}^{(p)𝒫lm}(\mathrm{\Omega }^{})^{}.$$ (A3) On $`S^3`$ the label $`p=3i,4i,..`$. It is related to the usual angular momentum $`k`$ by $`p=i(k+1)`$. The ranges of the other labels is then $`0lk`$ and $`lml`$. On $`H^3`$ there is a continuum of eigenvalues $`p[0,\mathrm{})`$. We will assume from now that the eigenmodes on are normalised by the condition $$\sqrt{\gamma }d^3xq_{𝒫lm}^{(p)ij}q_{𝒫^{}l^{}m^{}ij}^{(p^{})}=\delta ^{pp^{}}\delta _{𝒫𝒫^{}}\delta _{ll^{}}\delta _{mm^{}}$$ (A4) The bitensor $`W_{(p)i^{}j^{}}^{ij}(\mu )`$ appearing in our Green function has some additional properties arising from its construction in terms of the transverse and traceless tensor harmonics $`q_{ij}^{(p)𝒫lm}`$. The tracelessness of $`W_{iji^{}j^{}}^{(p)}`$ allows one to eliminate two of the coefficient functions in (A2). It may then be written as $`W_{iji^{}j^{}}^{(p)}(\mu )`$ $`=`$ $`w_1^{(p)}\left[g_{ij}^{}3ni^{}n_j^{}\right]\left[g_{i^{}j^{}}^{}n_i^{}^{}n_j^{}^{}\right]+w_2^{(p)}\left[4n_{(i}^{}g_{j)(i^{}}^{}n_{j^{})}^{}+4n_i^{}n_j^{}n_i^{}^{}n_j^{}^{}\right]`$ (A6) $`+w_3^{(p)}\left[g_{ii^{}}^{}g_{jj^{}}^{}+g_{ji^{}}^{}g_{ij^{}}^{}2n_i^{}g_{i^{}j^{}}^{}n_j^{}2n_i^{}^{}g_{ij}^{}n_j^{}^{}+6n_i^{}n_j^{}n_i^{}^{}n_j^{}^{}\right]`$ This expression is traceless on either index pair $`ij`$ or $`i^{}j^{}`$. The requirement that the bitensor be transverse $`^iW_{iji^{}j^{}}^{(p)}=0`$ and the eigenvalue condition $`(\mathrm{\Delta }_3\lambda _p)W_{(p)}^{iji^{}j^{}}=0`$ impose additional constraints on the remaining coefficient functions $`w_j^{(p)}(\mu )`$. To solve these constraint equations it is convenient to introduce the new variables on $`S^3`$ (on $`H^3`$, $`\mu `$ is replaced by $`i\stackrel{~}{\mu }`$) $$\{\begin{array}{ccc}\alpha (\mu )\hfill & =\hfill & w_1^{(p)}(\mu )+w_3^{(p)}(\mu )\hfill \\ \beta (\mu )\hfill & =\hfill & \frac{7}{(p^2+9)\mathrm{sin}\mu }\frac{d\alpha (\mu )}{d\mu }\hfill \end{array}$$ (A7) In terms of a new argument $`z=\mathrm{cos}^2(\mu /2)`$ (or its continuation on $`H^3`$) the transversality and eigenvalue conditions imply for $`\alpha (z)`$ $$z(1z)\frac{d^2\alpha (z)}{d^2z}+\left[\frac{7}{2}7z\right]\frac{d\alpha (z)}{dz}=(p^2+9)\alpha (z)$$ (A8) and then for the coefficient functions $`\{\begin{array}{ccc}w_1\hfill & =\hfill & Q_p\left(\left[2(\lambda _p6)z(z1)2\right]\alpha (z)+\frac{4}{7}\left[(\lambda _p+6)z(z\frac{1}{2})(z1)\right]\beta (z)\right)\hfill \\ w_2\hfill & =\hfill & Q_p\left(2(1z)\left[(\lambda _p6)z+3\right]\alpha (z)\frac{4}{7}\left[(\lambda _p+6)z(z1)(z\frac{3}{2})\right]\beta (z)\right)\hfill \\ w_3\hfill & =\hfill & Q_p\left(\left[2(\lambda _p6)z(z1)+3\right]\alpha (z)\frac{4}{7}\left[(\lambda _p+6)z(z\frac{1}{2})(z1)\right]\beta (z)\right)\hfill \end{array}`$ (A12) with $`\lambda _p=(p^2+3)`$. The above conditions leave the overall normalisation of the bitensor undetermined. To fix the normalisation constant $`Q_p`$ we contract the indices in the coincident limit $`z1`$. This yields $$W_{ij}^{(p)ij}(\mathrm{\Omega },\mathrm{\Omega })=\underset{𝒫lm}{}q_{ij}^{(p)𝒫lm}(\mathrm{\Omega })q^{(p)𝒫lmij}(\mathrm{\Omega })^{}=30Q_p\alpha (1).$$ (A13) By integrating over the three-sphere and using the normalisation condition (A4) on the tensor harmonics one obtains $`Q_p=\frac{p^2+4}{30\pi ^2\alpha (1)}`$. Notice that (A8) is precisely the hypergeometric differential equation, which has a pair of independent solutions $`\alpha (z)=_2F_1(3+ip,3ip,7/2,z)`$ and $`\alpha (1z)=_2F_1(3+ip,3ip,7/2,1z)`$. The former of these solutions is singular at $`z=1`$, i.e. for coincident points on the three-sphere, and the latter is singular for opposite points. The solution for $`\beta (z)`$ follows from (A7) and is given by $$\beta (z)=_2F_1(4ip,4+ip,9/2,z).$$ (A14) The hypergeometric functions are related by the transformation formula (eq.\[15.3.6\] in ) $`{}_{2}{}^{}F_{1}^{}(a,b,c,z)={\displaystyle \frac{\mathrm{\Gamma }(c)\mathrm{\Gamma }(cab)}{\mathrm{\Gamma }(ca)\mathrm{\Gamma }(cb)}}_2F_1(a,b,a+bc,1z)`$ (A15) $`+{\displaystyle \frac{\mathrm{\Gamma }(c)\mathrm{\Gamma }(a+bc)}{\mathrm{\Gamma }(a)\mathrm{\Gamma }(b)}}(1z)_2^{cab}F_1(ca,cb,cab,1z).`$ (A16) Only for the eigenvalues of the Laplacian on $`S^3`$, i.e. $`p=in(n3)`$, the term on the second line vanishes for $`{}_{2}{}^{}F_{1}^{}(3+ip,3ip,7/2,z)`$. For these special values, $`\alpha (z)`$ and $`\alpha (1z)`$ are no longer linearly independent but related by a factor of $`(1)^{n+1}`$, and they are both regular for any angle on the three-sphere. In fact, the hypergeometric series terminates for these parameter values and the hypergeometric functions reduce to Gegenbauer polynomials $`C_{n3}^{(3)}(12z)`$. We have a choice between using $`\alpha (z)`$ and $`\alpha (1z)`$ in the bitensor for these values of $`p`$. Since $`F(1z)1`$ for coincident points, it is more natural to choose $`\alpha (1z)`$ in the bitensor appearing in the Euclidean Green function (19). However, to obtain the Lorentzian correlator, we had to express the discrete sum (19) as a contour integral. Since the Euclidean correlator obeys a differential equation with a delta function source at $`\mu =0`$, we must maintain regularity of the integrand at $`\mu =\pi `$ when extending the bitensor in the complex $`p`$-plane. In other words, for generic $`p`$, we need to work with the solution $`\alpha (z)`$, rather than $`\alpha (1z)`$. Therefore, in order to write the Euclidean correlator as a contour integral, we first have replaced $`F(1z)`$ by $`F(z)(1)^{n+1}`$, by applying (A15) to (19), and we then have continued the latter term to $`(\mathrm{cosh}p\pi )_2^1F_1(3+ip,3ip,\frac{7}{2},z)`$. We conclude that the properties of the bitensor appearing in the tensor correlator completely determine its form. Notice that in terms of the label $`p`$ we have obtained a ’unified’ functional description of the bitensor $`W_{(p)}^{iji^{}j^{}}`$ on $`S^3`$ and $`H^3`$. Its explicit form is very different in both cases however, because the label $`p`$ takes on different values. But it is precisely this description that has enabled us in Section IV to analytically continue the correlator from the Euclidean instanton into open de Sitter space without Fourier decomposing it. We shall conclude this Appendix by giving the explicit formulae for the coefficient functions of the bitensor $`W_{iji^{}j^{}}^{L(p)}`$ appearing in our final result (36). With this description, they can be obtained by analytic continuation from $`S^3`$. To perform the continuation to $`H^3`$ we note that the geodesic separation $`\mu `$ on $`S^3`$ continues to $`i\chi `$ where $`\chi `$ is the comoving separation on $`H^3`$. Hence the hypergeometric functions on $`H^3`$ are defined by analytic continuation (eq. 15.3.7 in ) and may be expressed in terms of associated Legendre functions as $`\{\begin{array}{cc}\alpha (z)\hfill & =15\sqrt{\frac{\pi }{2}}(\mathrm{sinh}\chi )^{5/2}P_{1/2+ip}^{5/2}(\mathrm{cosh}\chi ),\hfill \\ \beta (z)\hfill & =15\sqrt{\frac{\pi }{2}}(\mathrm{sinh}\chi )^{7/2}P_{1/2+ip}^{7/2}(\mathrm{cosh}\chi ).\hfill \end{array}`$ (A19) Using the relation $`\mathrm{cosh}(\chi )=\mathrm{cosh}(\chi i\pi )`$, the Legendre functions on $`H^3`$ may be expressed as $`\{\begin{array}{ccc}P_{1/2+ip}^{5/2}(\mathrm{cosh}\chi )\hfill & =\hfill & \sqrt{\frac{2}{\pi \mathrm{sinh}\chi }}(1+p^2)^1(4+p^2)^1[3\mathrm{coth}\chi \mathrm{cosh}p(\pi +i\chi )\hfill \\ & & \frac{i\mathrm{sinh}p(i\chi +\pi )}{2p}((2p^2)(1+\mathrm{coth}^2\chi )+(4+p^2)\mathrm{cosech}^2\chi )]\hfill \\ P_{1/2+ip}^{7/2}(\mathrm{cosh}\chi )\hfill & =\hfill & \sqrt{\frac{2}{\pi \mathrm{sinh}\chi }}(1+p^2)^1(4+p^2)^1(9+p^2)^1\times \hfill \\ & & [\mathrm{cosh}p(\pi +i\chi )(p^21115\mathrm{c}\mathrm{o}\mathrm{s}\mathrm{e}\mathrm{c}\mathrm{h}^2\chi )\hfill \\ & & 6\frac{i\mathrm{sinh}p(i\chi +\pi )}{p}((1p^2)\mathrm{coth}^3\chi +(p^2+\frac{3}{2})\mathrm{coth}\chi \mathrm{cosech}^2\chi )]\hfill \end{array}`$ (A25) In the text, we have extracted the factors $`e^{\pm p\pi }`$ in these expressions in order to make contact with the usual description of the tensor correlator in terms of tensor harmonics on $`H^3`$. The coefficient functions of the bitensor $`W_{iji^{}j^{}}^{L(p)}(\chi )`$ in our final result (36) for the tensor correlator are $`\{\begin{array}{ccc}w_1\hfill & =\hfill & \frac{\mathrm{cosech}^5\chi }{4\pi ^2(p^2+1)}[\frac{\mathrm{sin}p\chi }{p}(3+(p^2+4)\mathrm{sinh}^2\chi p^2(p^2+1)\mathrm{sinh}^4\chi )\hfill \\ & & \mathrm{cos}p\chi (3/2+(p^2+1)\mathrm{sinh}^2\chi )\mathrm{sinh}2\chi ]\hfill \\ w_2\hfill & =\hfill & \frac{\mathrm{cosech}^5\chi }{4\pi ^2(p^2+1)}[\frac{\mathrm{sin}p\chi }{p}(3+12\mathrm{cosh}\chi 3p^2(1+2\mathrm{cosh}\chi )\mathrm{sinh}^2\chi \hfill \\ & & +p^2(p^2+1)\mathrm{sinh}^4\chi )+\mathrm{cos}p\chi (123\mathrm{cosh}\chi \hfill \\ & & +2(p^22)\mathrm{sinh}^2\chi +2(p^2+1)\mathrm{cosh}\chi \mathrm{sinh}^2\chi )\mathrm{sinh}\chi ]\hfill \\ w_3\hfill & =\hfill & \frac{\mathrm{cosech}^5\chi }{4\pi ^2(p^2+1)}[\frac{\mathrm{sin}p\chi }{p}(33p^2\mathrm{sinh}^2\chi +p^2(p^2+1)\mathrm{sinh}^4\chi )\hfill \\ & & +\mathrm{cos}p\chi (3/2+(p^2+1)\mathrm{sinh}^2\chi )\mathrm{sinh}2\chi ]\hfill \end{array}`$ (A33) As mentioned before, the bitensor $`W_{iji^{}j^{}}^{L(p)}`$ equals the sum (A3) of the rank-two tensor eigenmodes with eigenvalue $`\lambda _p=(p^2+3)`$ of the Laplacian on $`H^3`$. For $`\chi 0`$ these functions converge and they exponentially decay at large geodesic distances.
warning/0003/hep-ph0003050.html
ar5iv
text
# Vacuum structure of a modified MIT Bag ## Acknowledgments We thank K. Kirsten for a stimulating correspondence. Financial contributions from the Foundation for Fundamental Research (M.S. and R.D.V.), the Deutsches Bundesministerium fuer Bildung und Forschung, contract No. 06 Tue 887, (T.G.), and the Graduiertenkolleg “Struktur und Wechselwirkung von Hadronen und Kernen” (R.H.), are gratefully acknowledged.
warning/0003/astro-ph0003478.html
ar5iv
text
# Evolutionary Timescale of the DAV G117-B15A: The Most Stable Optical Clock Known ## 1 Introduction We report our continuing study of the star G 117–B15A, also called RY LMi, and WD0921+354, one of the hottest of the pulsating white dwarfs with hydrogen atmospheres, the DAV or ZZ Ceti stars (McGraw 1979). McGraw & Robinson (1976) found the star was variable, and Kepler et al. (1982) studied its light curve and found 6 pulsation modes. The dominant mode is at 215 s, has a fractional amplitude of 22 mma, and is stable in amplitude and phase. The other, smaller pulsation modes vary in amplitude from night to night (Kepler et al. 1995). Because the DAVs appear to be normal stars except for their variability (Robinson 1979, Bergeron et al. 1995), it is likely that the DAV structural properties are representative of all DA white dwarfs. The rate of change of a pulsation period is directly related to the evolutionary timescale of a white dwarf, allowing us to directly infer the age of a white dwarf since its formation. We have been working since 1975 to measure the rate of period change with time ($`\dot{P}`$) for the $`P=215`$ s periodicity of G117–B15A, and the Kepler et al. (1991) determination was $`\dot{P}=(12.0\pm 3.5)\times 10^{15}\mathrm{s}/\mathrm{s}`$, including all data obtained from 1975 through 1990. Kepler (1984) demonstrated that the observed variations in the light curve of G 117–B15A are due to non-radial g-mode pulsations and therefore the timescale for period change is directly proportional to the cooling timescale. For comparison, the most stable atomic clocks have rates of period change of the order of $`\dot{P}2\times 10^{14}`$ s/s, while the most precise millisecond pulsars have $`\dot{P}10^{20}`$ s/s (Kaspi, Taylor & Ryba 1994 and references therein). Since the stability of a clock is measured by $`P/\dot{P}`$, G117–B15A has the same order of stability as the most stable millisecond pulsar. G117–B15A is the first pulsating white dwarf to have its main pulsation mode index identified. The 215 s mode is an $`\mathrm{}=1`$, as determined by comparing the ultraviolet pulsation amplitude (measured with the Hubble Space Telescope) to the optical amplitude (Robinson et al. 1995). Robinson et al. (1995), and Koester, Allard & Vauclair (1994) derive $`T_{\mathrm{eff}}`$ near 12,400 K, while Bergeron et al. (1995), using a less efficient model for convection, derives $`T_{\mathrm{eff}}`$=11,600 K. Bradley (1996) used the mode identification and the observed periods of the 3 largest known pulsation modes to derive a hydrogen layer mass lower limit of $`10^6M_{}`$, and a best estimate of $`1.5\times 10^4M_{}`$, assuming $`k=2`$ for the 215 s mode, and 20:80 C/O core mass. The core composition is constrained mainly by the presence of the small 304 s pulsation. ## 2 Observations We obtained 19.6 h of time series photometry in Dec 1996 and Feb 1997, plus 18.8 h in Mar and Dec 1999, using the three-star (Kleinman, Nather & Phillips 1996) photometer on the $`2.1`$ m Struve telescope at McDonald Observatory. To maximize the signal-to-noise (S/N) we observed unfiltered light, because the nonradial g-mode light variations have the same phase in all colors (Robinson, Kepler & Nather 1982). G117–B15A has V=15.52 (Eggen & Greenstein 1965). ## 3 Data Reduction We reduce and analyze the data in the manner described by Nather et al. (1990), and Kepler (1993). We bring all the data to the same fractional amplitude scale, and transform the observatories’ UTC times to the uniform Barycentric Coordinate Time (TCB) scale (Standish 1998), using JPL DE96 ephemeris as our basic solar system model (Stumpff 1980). Kaspi, Taylor & Ryba (1994) show that the effects of using different atomic timescales and ephemeris are negligible. We compute Fourier transforms for each individual run, and verify that the main pulsation at 215 s dominates each data set and has a stable amplitude. ## 4 Time Scale for Period Change As the dominant pulsation mode at P=215 s has a stable frequency and amplitude since our first observations in 1975, we can calculate the time of maximum for each new run and look for deviations due to evolutionary cooling. We fit our observed time of maximum light to the equation: $$(OC)=\mathrm{\Delta }E_0+\mathrm{\Delta }PE+\frac{1}{2}P\dot{P}E^2,$$ where $`\mathrm{\Delta }E_0=(T_{max}^0T_{max}^1)`$, $`\mathrm{\Delta }P=(PP_{t=T_{max}^0})`$, and $`E`$ is the epoch of the time of maximum, i.e, the number of cycles after our first observation. In Figure 1, we show the O–C timings after subtracting the correction to period and epoch, and our best fit curve through the data. ¿From our data through 1999, we obtain a new value for the epoch of maximum, $`T_{max}^0=\mathrm{244\hspace{0.17em}2397.917509}\mathrm{TCB}\pm 0.5\mathrm{s}`$, a new value for the period, $`P=215.1973907\pm 0.0000006\mathrm{s}`$, and most importantly, a rate of period change of: $$\dot{P}=(2.3\pm 1.4)\times 10^{15}\mathrm{s}/\mathrm{s}.$$ We use linear least squares to make our fit, with each point weighted inversely proportional to the uncertainty in the time of maxima for each individual run squared. We quadratically add an additional $`1.8`$ s of uncertainty to the time of maxima for each night to account for external uncertainty caused perhaps by the beating of small amplitude pulsations (Kepler et al. 1995) or small amplitude modulation. The estimated $`\dot{P}`$ is substantially different from the value estimated in 1991. The apparent reason is a scatter of the order of 1.8 s present in the measured times of maxima. Kepler et al. (1995) discuss the possibility of such scatter being caused by modulation due to nearby frequencies, and Costa et al. (1999) shows that the real uncertainties must include the effect of all periodicities present. The 1991 value did not include such scatter in the uncertainty estimation, and resulted in an overestimated statistical accuracy. We now treat this scatter as an external source of noise. ## 5 Core Composition For a given mass and internal temperature distribution, theoretical models show that the rate of period change increases if the mean atomic weight of the core is increased, for models which have not yet crystallized in their interiors. This applies to G117–B15A, as it is not cool enough to have a crystallized core (Winget et al. 1997). Bradley, Winget & Wood (1992) and Bradley (1998) compute rates of period change for models that are applicable to G117–B15A, and we summarize the relevant results here. The models of Bradley, Winget, & Wood (1992) and Bradley (1998) are full evolutionary models that include compositional stratification, accurate physics, and use the most recent neutrino emission rates. We refer the reader to Bradley, Winget, & Wood (1992) and Bradley (1996, 1998) for further details. Two major known processes govern the rate of period change in the theoretical models of the ZZ Ceti stars: residual gravitational contraction, which causes the periods to become shorter, and cooling of the star, which increases the period as a result of the increasing degeneracy (Winget, Hansen, & Van Horn 1983), given by $$\frac{d(\mathrm{ln}P}{dt}=a\frac{d\mathrm{ln}T_c}{dt}+b\frac{d\mathrm{ln}R}{dt}$$ where $`a`$ and $`b`$ are constants associated with the rate of cooling and contraction respectively, and are of order unity. Following Kawaler, Hansen, & Winget (1985), we can write $$\frac{d\mathrm{ln}P}{dt}=(a+bs)\frac{d\mathrm{ln}T_c}{dt}$$ where $`s`$ is the ratio of the contraction rate to the cooling rate $$s\frac{d\mathrm{ln}T_c}{dt}=\frac{d\mathrm{ln}R}{dt}$$ or $$s=\frac{d\mathrm{ln}R}{d\mathrm{ln}T_c}.$$ The $`dt`$ terms cancel because we evaluate the derivative as the differences in the radius, core temperature, and age between two models. Spectroscopic $`\mathrm{log}g`$ values suggest that G117-B15A has a mass between 0.53 $`M_{}`$ (Koester & Allard 2000) and 0.59 $`M_{}`$ (Bergeron et al. 1995), and this agrees with the preferred seismological mass range of 0.55 to 0.60 $`M_{}`$ (Bradley 1998). For a DA white dwarf near 12,000 K, the radius is about $`9.6\times 10^8`$ cm, with a contraction rate of about 1 cm yr<sup>-1</sup>. The core temperature is about $`1.2\times 10^7`$ K, with a cooling rate of about $`0.05`$ K yr<sup>-1</sup>. With these numbers, $`s`$ is about $`0.025`$, which confirms our expectation that the rate of period change is dominated by cooling. Other processes, such as rotational spin-down and magnetic fields must be small, because we do not see reliable evidence of either in the fine structure splitting of the observed frequencies. Bradley (1998) give a $`\dot{P}`$ value of $`3.7\times 10^{15}\mathrm{s}/\mathrm{s},`$ and find a spread of $`\pm 1\times 10^{15}\mathrm{s}/\mathrm{s},`$ predicted by the range of acceptable models for G117–B15A, with the $`0.60M_{}`$ models having the smaller values. His predicted value is within the $`1\sigma `$ error bars of the observed value; a more precise observational $`\dot{P}`$ determination could in principle suggest a favored stellar mass. Bradley’s (1998) models are typically about 80% oxygen, and Bradley et al. (1992) describe in detail the effect of changing the core composition from pure carbon to pure oxygen for 0.5 and 0.60 $`M_{}`$ models. They also show that the predicted $`\dot{P}`$ value from an oxygen core model is about 15 to 20% larger than for an equivalent carbon core model, rather than the 33% predicted by Mestel (1952) cooling theory. This reduction in $`\dot{P}`$ from Mestel theory is the result of the ions being a Coulomb liquid, rather than an ideal gas as assumed by Mestel theory. The $`\dot{P}`$ values quoted above are for the case where the 215 s mode is not trapped (see Bradley 1996 for details), and Bradley et al. (1992) show that if the 215 s mode is trapped, then the predicted $`\dot{P}`$ value could be as little as half the values predicted by Bradley (1998) and quoted above. In recent years, the $`\dot{P}`$ determinations have fluctuated between about 1 and $`3\times 10^{15}\mathrm{s}/\mathrm{s}`$ (see Table 1), so the values predicted by seismological models are still consistent with the observations. Reducing the observational errors to about half the present value of $`1.4\times 10^{15}\mathrm{s}/\mathrm{s}`$ would provide enough of a constraint to confront the model predictions. ## 6 Reflex Motion The presence of an orbital companion could contribute to the period change we have detected. When a star has an orbital companion, the variation of its line-of-sight position with time produces a variation in the time of arrival of the pulsation maxima, by changing the light travel time between the star and the observer. Kepler et al. (1991) calculated the possible contribution to $`\dot{P}`$ caused by reflex orbital motion of the observed proper motion companion of G117–B15A as $`\dot{P}1.9\times 10^{15}\mathrm{s}/\mathrm{s}`$. If the orbit is highly eccentric and G117–B15A is near periastron, the orbital velocity could not be higher than twice that derived above or it would exceed escape velocity. The above derivation assumed that the orbit is nearly edge on to give the largest effect possible. Therefore, $`\dot{P}_{orb}3.8\times 10^{15}`$ s/s. The upper limit to the rate of period change could also be expected if a planet of Jupiter’s mass were orbiting the WD at a distance of 24 AU, which corresponds to an orbital period of 118 yr, or a smaller planet in a closer orbit. Note that reflex motion produces sinusoidal variations on the $`OC`$, which are only distinguishable from parabolic variations after a significant portion of the orbit has been covered. As we have observed the star for 25 yr, a sinusoid with a period shorter than 100 yr can be discarded, but if the orbiting object were near apoastron in a highly eccentric orbit, the difference would be harder to distinguish. ## 7 Proper Motion Pajdosz (1995) discusses the influence of the proper motion of the star on the measured $`\dot{P}`$: $$\dot{P}_{\mathrm{obs}}=\dot{P}_{\mathrm{evol}}\left(1+v_r/c\right)+P\dot{v}_r/c$$ where $`v_r`$ is the radial velocity of the star. Assuming $`v_r/c1`$, he derived $$\dot{P}_{\mathrm{pm}}=2.430\times 10^{18}P[s]\left(\mu [\mathrm{"}/yr]\right)^2\left(\pi [\mathrm{"}]\right)^1$$ where $`\dot{P}_{\mathrm{pm}}`$ is the effect of the proper motion on the rate of period change, $`P`$ is the pulsation period, $`\mu `$ is the proper motion, and $`\pi `$ is the parallax. He also calculated that for G117–B15A, $`\dot{P}_{\mathrm{pm}}(8.0\pm 0.4)\times 10^{16}`$ s/s, using the proper motion ($`\mu =0.136\pm 0.002\mathrm{"}/\mathrm{yr}`$) and parallax $`\pi =(0.012\pm 0.005\mathrm{"})`$ measured by Harrington & Dahn (1980). With the parallax by Van Altena et al. (1995) of $`\pi =(0.0105\pm 0.004\mathrm{"})`$, and the above proper motion, we calculate $`\dot{P}_{\mathrm{pm}}=(9.2\pm 0.5)\times 10^{16}`$ s/s. The upper limit to the observed $`\dot{P}`$ is already only a few times the $`\dot{P}`$ expected from proper motion alone. ## 8 Conclusions While it is true that the period change timescale can be proportional to the cooling timescale, other phenomena with shorter timescales can affect $`\dot{P}`$. The cooling timescale is the longest possible one. As a corollary, if the observed $`\dot{P}`$ is low enough to be consistent with evolution, then other processes (such as perhaps a magnetic field) are not present at a level sufficient to affect $`\dot{P}`$. We compare the observed value of $`\dot{P}`$ with the range of theoretical values derived from realistic evolutionary models with $`C/O`$ cores subject to g–mode pulsations in the temperature range of G117–B15A. The adiabatic pulsation calculations of Bradley (1996), and Brassard et al. (1992,1993), which allow for mode trapping, give $`\dot{P}`$$`(27)\times 10^{15}\mathrm{s}/\mathrm{s}`$ for the $`\mathrm{}=1`$, low k oscillation observed. The observed 3$`\sigma `$ upper limit, $`\dot{P}6.5\times 10^{15}\mathrm{s}/\mathrm{s}`$, corresponding to a timescale for period change of $`P/\dot{P}1.2\times 10^9`$ yr, equivalent to 1 s in $`6\times 10^6`$ yr, is within the theoretical predictions and very close to it. Our upper limit to the rate of period change brings us to realms where reflex motion from the proper motion companion, if they form a physical binary, or an unseen orbiting planet is of the same order as the evolutionary timescale. The effect of proper motion of the star itself is only a few times smaller. These two effects must therefore be accurately measured. We are on the way to measure the evolutionary time scale for this lukewarm white dwarf, but the observed phase scatter of the order of 1.8 s increased the baseline necessary for a measurement. This scatter is still present in our measurement. This work was partially supported by grants from CNPq (Brazil), FINEP (Brazil), NSF (USA), NASA (USA).
warning/0003/cond-mat0003519.html
ar5iv
text
# Heterogeneous condensation in dense media ## 1 Profile around the solitary droplet Due to the external influence in the initial moment of time one can observe in the system the homogeneous mother metastable phase with the particle number density $`n`$ which is equal to some value $`n_0`$. All heterogeneous centers are distributed rather homogeneously in space with the number density $`\eta _{tot}`$. The process of condensation can begin only when $`n_0`$ is greater than the molecule number density $`n_{\mathrm{}}`$ in the vapor saturated over the plane liquid. The power of metastability of vapor will be characterized by the value of a supersaturation $`\zeta `$ defined as $$\zeta =\frac{n}{n_{\mathrm{}}}1$$ The initial value of the supersaturation will be marked by $`\zeta _0`$. Practically immediately around every center there will be formed an equilibrium embryo which has $`\nu _e`$ molecules of the condensing substance. The value of $`\nu _e`$ is relatively small<sup>5</sup><sup>5</sup>5In comparison with the characteristic number of molecules inside the droplet during the nucleation period. and there is no need to consider the density profile around the equilibrium embryos<sup>6</sup><sup>6</sup>6In fact it will disappear rather fast. This leads to the slight variation of the equilibrium embryo characteristics. This variation will act on the gap. But the final relaxation will be rather rapid.. The number of the equilibrium embryos in the unit volume<sup>7</sup><sup>7</sup>7The system of a unit volume will be considered. will be equal to $`\eta _{tot}`$. During the condensation process the number of the equilibrium embryos $`\eta `$ will be decreased due to the exhaustion of the free heterogeneous centers $$\eta =\eta _{tot}N$$ where $`N`$ is number of the supercritical embryos which will be called as the droplets. The last equation illustrates the main specific feature in kinetics of heterogeneous condensation. Despite the simple form of the last relation the effect is very complicate because $`N`$ depends on time in a very complicate manner. The effects of the density profile will be essential also for $`N`$ and one can not directly apply the results of . One has to determine the effect of the influence on $`N`$ even for the density profile of the solitary droplet. We shall call an approach where the law of embryos growth is found from continuous model but there is no account of the profile around the droplets as the ”Additive approach” (AA). Then one can formulate the evident Statement 1 The duration of the nucleation period<sup>8</sup><sup>8</sup>8The period of nucleation is the period of the relatively intensive formation of the droplets. It can be proven that the end of this period is well defined due to the cut-off of the intensity of the droplets formation. and the characteristic sizes of the droplets at the end of nucleation period are greater than those calculated in AA Really, the existence of the density profile means that the part of substance is going to be consumed from the regions where there is already no droplets formation. This substance is consumed from the gap instead of unexhausted regions as it is done in AA. Then having repeated all constructions<sup>9</sup><sup>9</sup>9In the AA was formulated for external conditions of dynamic type. For the situation of decay the required hierarchical inequalities can be proven by the same way. One can note that in there is no special reference on the type of condition when the required estimates are proven. from one can see<sup>10</sup><sup>10</sup>10The barrier character of nucleation is required here. It means that the magnitude of the activation barrier height has the same order as the free energy of the homogeneous critical embryos (may be it is $`4`$ or $`3`$ times smaller). Statement 2 The characteristic size of the droplets at the end of the nucleation period is many times greater than the size of the critical embryo. The main role in vapor consumption is played by the supercritical embryos. and Statement 3 The characteristic time of the nucleation period duration is many times greater than the time of relaxation to the stationary state in the nearcritical region. Then one can use the stationary rate of nucleation as the intensity of the droplets formation in every current moment of time. Due to the statement 2 one has to investigate the profile around a growing droplet. The problem is whether one has to consider the interference of profiles around different droplets. To solve this problem one has to use the small parameter of the theory. Due to the statement 3 the rate of nucleation is equal to the stationary one. It can be taken from $$I_s=Z\eta exp(\mathrm{\Delta }F)$$ where $`\mathrm{\Delta }F`$ is the height of activation barrier (taken in thermal units), $`\eta `$ is the number of free (unoccupied by the supercritical embryos) heterogeneous centers, $`Z`$ is Zeldowitch factor. Zeldowitch factor is the smooth function of the supersaturation which is given by $$Z=\frac{W}{\pi ^{1/2}\mathrm{\Delta }\nu _e\mathrm{\Delta }\nu _c}$$ where $`W`$ is kinetic factor, $`\mathrm{\Delta }\nu _c`$ is the halfwidht of the nearcritical region, $`\mathrm{\Delta }\nu _e`$ is the width of the equilibrium region. Due to the rather small size of the critical embryo it is reasonable to use the free molecular regime of the substance exchange<sup>11</sup><sup>11</sup>11As long as the characteristic size of the droplet during the nucleation is many times greater that the critical size the diffusion regime of growth for the characteristic droplets is quite reasonable.. Namely, in this regime an expression for the nucleation rate is well based. One has also to note that the critical embryo is in equilibrium (unstable one) with the metastable phase which implies no profiles of vapor density and regime of the substance exchange is the free molecular one. Under the free molecular regime $`W`$ can be calculated as $$W=3\frac{\zeta +1}{\tau }\nu _c^{2/3}\alpha $$ where $`\nu _c`$ is the number of molecules inside the critical embryo, $`\alpha `$ is the condensation coefficient, $$\tau =12[(36\pi v_l^2)^{1/3}n_{\mathrm{}}v_T]^1$$ is the characteristic time, $`v_l`$ is a volume per one molecule in a liquid phase, $`v_T`$ is the mean thermal velocity of a molecule. The value of $`\mathrm{\Delta }\nu _c`$ is the halfwidth of the nearcritical region and it can be calculated as $$\mathrm{\Delta }\nu _c=\underset{\nu (\nu _c+\nu _e)/2}{}\mathrm{exp}(F_c+F_\nu )\pi ^{1/2}$$ where $`\nu `$ is the number of the molecules inside the embryo, $`F_\nu `$ is the free energy of the embryo of $`\nu `$ molecules, $`F_c`$ is the free energy of the critical embryo. In continuous approximation it can be calculated as<sup>12</sup><sup>12</sup>12Ordinary $`\mathrm{\Delta }\nu _e`$ is smaller than $`\mathrm{\Delta }\nu _c`$ and an explicit summation for $`\mathrm{\Delta }\nu _c`$ is quite reasonable. $$\mathrm{\Delta }\nu _c=|\frac{2}{\frac{\delta ^2F}{\delta \nu ^2}}|_{\nu =\nu _c}|^{1/2}$$ The value of $`\mathrm{\Delta }\nu _e`$ can be calculated as $$\mathrm{\Delta }\nu _e=\underset{\nu (\nu _e+\nu _c)/2}{}\mathrm{exp}(F_\nu +F_e)$$ where $`F_e`$ is the free energy of the equilibrium embryo. Both $`\mathrm{\Delta }\nu _c`$ and $`\mathrm{\Delta }\nu _e`$ are rather smooth functions of the supersaturation. One can see that $`I_s`$ is a very sharp function of the supersaturation. It means that the relatively small fall of the supersaturation leads to the interruption of the droplets formation. At least for all $`\zeta >\zeta _0/2`$ one can show that $`d^2\zeta /dt^2>0`$ and the is no long tail of a size spectrum with a small intensity. It means that the interruption of the droplets formation leads to the interruption of the nucleation process. So, the relative fall of the supersaturation during the nucleation process is small. One can come to the following statement Statement 4 During the nucleation period the relative variation of supersaturation is small. The last statement shows that there is no need to consider the interference of profiles in order to change the rate of droplets growth. On the base of mentioned expressions and the smallness of the relative fall of supersaturation one can see the validity of the following approximation $$I_s(\zeta )=I_s(\zeta _0)\mathrm{exp}(\mathrm{\Delta }F(\zeta _0)\mathrm{\Delta }F(\zeta ))$$ for the nucleation period. Moreover one can linearize the height of the activation barrier over the supersaturation and get $$I_s(\zeta )=I_s(\zeta _0)\mathrm{exp}(\frac{d\mathrm{\Delta }F(\zeta )}{d\zeta }|_{\zeta =\zeta _0}(\zeta \zeta _0))$$ (1) The validity of the last approximation depends on the concrete type of heterogeneous centers but for the majority of the heterogeneous centers types the last approximation is valid. For example, this validity can be directly proven for ions. One can explicitly calculate the derivative in the last expression<sup>13</sup><sup>13</sup>13Here we assume the vapor to be an ideal gas and suppose the possibility to present the free energy of critical and equilibrium embryos as an analytical function of an inverse embryo radius. $$\frac{d\mathrm{\Delta }F}{d\zeta }=\frac{1}{\zeta +1}(\nu _c\nu _e)$$ The smooth character of the last expression shows the validity of (1) one more time. Then (1) can be rewritten as $$I_s(\zeta )=I_s(\zeta _0)\mathrm{exp}(\mathrm{\Gamma }\frac{\zeta \zeta _0}{\zeta _0})$$ (2) where $$\mathrm{\Gamma }=\zeta _0\frac{d\mathrm{\Delta }F}{d\zeta }|_{\zeta =\zeta _0}=\frac{\zeta _0}{\zeta _0+1}(\nu _c(\zeta _0)\nu _e(\zeta _0))$$ As far as the value of $`\nu _c`$ in going in the presented theory to infinity the value of $`\mathrm{\Gamma }`$ is also going to $`\mathrm{}`$. The real value of $`\mathrm{\Gamma }`$ is very big. Certainly, one can consider the possibility of compensation between $`\nu _c`$ and $`\nu _e`$ in the expression for $`\mathrm{\Gamma }`$. Then one has to mention thatdue to the barrier character of nucleation at least $`\nu _c\nu _e\mathrm{\Delta }\nu _c`$. Having estimated $`\mathrm{\Delta }\nu _c`$ by the homogeneous value $`\mathrm{\Delta }\nu _c\nu _c^{2/3}`$ one can see that $`\mathrm{\Gamma }1`$ in any case. The small value of $`\mathrm{\Gamma }^1`$ will be very important in further constructions. We see that essential dependence over the supersaturation occurs through the height of activation barrier. It allows to give the interpretation of the stationary rate of nucleation as the probability for the given embryo to overcome the activation barrier. After the interpretation of $`I_s`$ as the probability we can apply it to the arbitrary spatial point of spatially unhomogeneous system. To use this interpretation the natural requirement is the weak unhomogenity of the system. Namely, the volume of the regions where $$\frac{\zeta (r)\zeta (r+\sqrt{4Dt_s})}{\zeta (r)}\mathrm{\Gamma }^1$$ is violated has to be relatively small. Here $`D`$ is the diffusion coefficient, $`t_s`$ is the time of relaxation in the nearcritical region which can be estimated according to Zeldowitch $$t_s\frac{\mathrm{\Delta }\nu _c^2}{W}$$ One can use instead of $`t_s`$ the time $`Z^1`$ which can be interpreted as the mean time to overcome the nearcritical region. Both these estimates are observed. Actually we need them only for those regions where the intensity of the droplets formation isn’t too small in comparison with initial intensity. Certainly, the required property is observed there. Now we have to turn to the rate of embryos growth. According to the Statement 2 the characteristic size of droplets is rather big. Then it is more reasonable to use the diffusion regime of droplets growth. Under the intermediate Knudsen numbers one has to use an interpolation law for the rate of embryos growth (for example, see ). It will be important that all expressions for the embryos growth lead to the avalanche character of the substance consumption. The avalanche character of the substance consumption means that the quantity of substance accumulated by a droplet strongly increases in time. The most evident manifestation of the avalanche consumption can be seen in the free molecular regime of the substance consumption. The most ”weak” effect of the avalanche consumption can be seen under the diffusion regime of the substance consumption. The force of the iterations convergence in is based on this property. The property of avalanche consumption will be extremely important in the further constructions also. That’s why we take the diffusion regime to have the worst situation and to grasp errors in all possible cases. In the diffusion regime of the vapor consumption the law of growth for a droplet (i.e. for the supercritical embryo) can be written in the following way $$\frac{d\nu }{dt}=\kappa \zeta \nu ^{1/3}$$ where $$\kappa =(\frac{2}{3})^{1/3}4\pi n_{\mathrm{}}D(\frac{v_l}{2\pi })^{1/3}$$ is some constant. The last expression is written in stationary approximation. The non-stationary effects were investigated in many publications in details and they are rather small ones. One can see that the rate of the droplet growth is proportional to $`\zeta `$. So, the rate of growth can be essentially changed only by the essential relative variation of $`\zeta `$. Then according to Statement 4 one can see that Statement 5 The rate of droplets growth during the nucleation period can be approximated as a constant one. The last statement is extremely important because is allows to analyze the profile of the density initiated by a solitary droplet. Now this case will be considered. The approximately constant value of the supersaturation allows to integrate the law of growth and get $$\nu (t)=\gamma t^{3/2}$$ where $$\gamma =(4\pi )^{3/2}(\frac{3v_l}{4\pi })^{1/2}(\frac{2\zeta n_{\mathrm{}}D}{3})^{3/2}$$ and $`t`$ is the duration of the irreversible growth for the given droplet. Consider the spherical system of coordinates with a center in the center of droplet. Diffusion equation will be written as $$\frac{n}{t}=D\mathrm{\Delta }n$$ where $`\mathrm{\Delta }`$ is the Laplace operator. Diffusion coefficient $`D`$ is supposed to be approximately constant (there is a lot of a passive gas and the density of a gas mixture is approximately constant). The boundary conditions are the following $$n|_{r=\mathrm{}}=n(\mathrm{})$$ $$n|_{r=R_d}=n_{\mathrm{}}$$ where $`R_d`$ is the radius of a droplet. The values $`n_{\mathrm{}}`$ and $`n(\mathrm{})`$ are known parameters. The variable $`r`$ is the distance until the center of the embryo. The stationary approximation is suitable for the rate of droplets growth. The errors are analyzed in and they are small. But the stationary solution can not give any reasonable result for the density far from the droplet. Really, the stationary solution is the following $$n(r)=n(\mathrm{})\frac{R_d}{r}(n(\mathrm{})n_{\mathrm{}})$$ (3) and has a very long tail. This tail leads to the infinite value of $$G=_0^{\mathrm{}}4\pi r^2(n(\mathrm{})n(r))𝑑r$$ which has to be the integral excess of the substance. This quantity has to be in the droplet. This contradiction shows that it is absolutely impossible to use the stationary approximation for the density profile around the droplet. One has to introduce another approach. One can see that if the first boundary condition will be changed by $$n|_{r=\mathrm{}}=n(\mathrm{})(1\mathrm{\Gamma }^1)$$ then the rate of embryos growth will be changed unessentially. But the level $`n(\mathrm{})(1\mathrm{\Gamma }^1)`$ is the level when the nucleation stops. So, one can see that during the nucleation period there is no interaction between droplets through the change of the growth rate. Certainly, two droplets can appear too close and act one upon another but the probability of such coincidence is too small. That is why one can come to the ”Principle of a separate growth of droplets during the nucleation period”. Now one has to prove that at the distances $`(5÷10)R_d`$ from the droplet one can observe the quasistationary profile. One has to note that $$v_l/v_v1$$ (4) where $`v_v`$ is the partial molecule volume in a vapor phase. Really, the last ratio is very small (for example, this ratio is $`0.001`$ for water in normal thermodynamic conditions). But contrary to $`\mathrm{\Gamma }^1`$ one can not consider it in all cases as zero. Now one can introduce a formal parameter $`l`$ which attains big values $$l1$$ but satisfies the following condition $$l^2\frac{v_l}{v_v}1$$ (5) Due to (4) it is possible to do. In the region $`rlR_d`$ the stationary profile is established after $$t_h=\frac{l^2R_d^2}{4D}$$ It is necessary to show that $$s\frac{R_d(t+t_h)R_d(t)}{R_d(t)}1$$ Really, $$s\frac{dR_d}{dt}\frac{t_h}{R_d}$$ and $$sl^2\frac{v_l}{v_v}$$ which is a small value according to (5). So, the stationary form of profile in the region $`r<R_dl`$ is proven. As far as $`\mathrm{\Gamma }1`$ and at least $`\mathrm{\Gamma }l`$ one can see that in the region $`r<lR_d`$ there is no formation of new droplets. Then this region isn’t interesting for the theory and one can only observe the region $`r>lR_d`$. The previous notation is rather important. Namely this property allows to use the model with a point source. Really, one can consider only the distances greater than $`lR_d`$. But at these distances the droplet can be interpreted as a point. Certainly, the point approximation of a droplet can not give an expression for the rate of droplets growth because the boundary condition at $`r=R_d`$ is absent. But the rate of growth is already known and can be used directly as a known function of time. Really, $$\frac{d\nu }{dt}=\lambda t^{1/2}$$ where $$\lambda =2^{5/2}\pi v_l^{1/2}\zeta ^{3/2}n_{\mathrm{}}^{3/2}D^{3/2}$$ The action of a point source of a vapor consumption can be described in a simple and suitable manner under the Green function formalism. The Green function for the diffusion equation can be written in a following form $$Gr=\mathrm{\Theta }(t)\frac{\mathrm{exp}(\frac{r^2}{4Dt})}{(4\pi Dt)^{3/2}}$$ Then one can get the density profile by a simple integration $$n(r)=n(\mathrm{})_0^t\frac{\lambda x^{1/2}}{(4\pi D(tx))^{3/2}}\mathrm{exp}(\frac{r^2}{4D(tx)})𝑑x$$ After evident transformations one can come to $$\frac{\zeta _0\zeta }{\zeta _0}=\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}f(\beta )$$ (6) where $$\beta =\frac{r}{\sqrt{4Dt}}$$ and $$f(\beta )=_\beta ^{\mathrm{}}(\frac{1}{\beta ^2}\frac{1}{x^2})^{1/2}\mathrm{exp}(x^2)𝑑x$$ It is important that the profile dependence on $`t`$ and $`r`$ is now going through $`\beta `$. Concrete form of $`f(\beta )`$ is drawn in fig. 1. One can get for $`f(\beta )`$ an expression through special functions $$f(\beta )=\frac{1}{2}\mathrm{\Gamma }(\frac{3}{2})\mathrm{exp}(\beta ^2)\mathrm{\Psi }(\frac{3}{2},\frac{3}{2};\beta ^2)$$ Here $`\mathrm{\Gamma }`$ is the Gamma-function , $`\mathrm{\Psi }`$ is the confluent hypergeometric function. One can get asymptotes for $`f(\beta )`$ at small and big values of $`\beta `$. At small values $$f(\beta )\frac{\sqrt{\pi }}{2}\frac{1}{\beta }$$ (7) which corresponds to the stationary solution (3). At big values of $`\beta `$ one can come to $$f(\beta )=\mathrm{exp}(\beta ^2)\frac{1}{2\beta ^3}_0^{\mathrm{}}x^{1/2}\mathrm{exp}(x)𝑑x\frac{\mathrm{exp}(\beta ^2)}{\beta ^3}$$ (8) One can see that this asymptote radically differs from the stationary solution. Namely this tail behavior ensures the convergence of the integral for $`G`$. Certainly, the Green function formalism ensures a precise value for $`G`$ which is introduced here as an external object. Now one can turn to construct some approximation for the nucleation rate around the growing droplet. One can see that according to (2) the behavior of the supersaturation is important when $`\zeta _0\zeta (2÷3)\zeta _0/\mathrm{\Gamma }`$. When $`\zeta _0\zeta (2÷3)\zeta _0/\mathrm{\Gamma }`$ the intensity of the droplets formation is negligibly small. From (1) one can see that $$I_s(\zeta (r))=I_s(\zeta _0)\mathrm{exp}(\mathrm{\Gamma }\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}f(\beta ))$$ Then one can extract a positive parameter $$\sigma \mathrm{\Gamma }^2\frac{v_l}{v_v}$$ which will be important in further constructions. Due to $`\mathrm{\Gamma }1`$ one can easily see that $$\sigma 1$$ The last condition isn’t necessary for further constructions, but it will be rather important for manifestation of profile effects in the nucleation process. The last condition is also the most doubtful one because $`v_l/v_v1`$ and one has the ration of two big parameters with generally unknown result. It is necessary to stress that condition $`v_l/v_v1`$ isn’t so strong as $`\mathrm{\Gamma }1`$. In frames of thermodynamic description $`\mathrm{\Gamma }1`$ is the main condition necessary to give the thermodynamic description and $`v_l/v_v1`$ is the supplementary condition which isn’t absolutely necessary but slightly simplifies the theory. In the situation of homogeneous condensation one has the ”hidden contradiction between the thermodynamic description and the relatively intensive nucleation”. Really, as far as in the homogeneous condensation $`\mathrm{\Delta }F=F_c\nu _c^{2/3}`$ the limit $`\nu \mathrm{}`$ means $`\mathrm{\Delta }F\mathrm{}`$ and the rate of nucleation goes to zero. So, there is the contradiction between thermodynamic limit in the critical embryo description and the observable<sup>14</sup><sup>14</sup>14Which isn’t too small. rate of nucleation. In the case of heterogeneous condensation there is no such contradiction if there are some active centers of condensation. Then the height of the activation barrier has no direct connection with a number of molecules inside the critical embryo. For example, the halfwidth of the nearcritical region estimated by a homogeneous value has a value $`\nu ^{2/3}`$ and goes to infinity when $`\nu \mathrm{}`$, but the free energy decreases at the boundary of the nearcritical region only by one thermal unit. So, in certain sense the case of heterogeneous condensation is more preferable for theoretical description. As a compensation for this advantage one has to note that the Statement 1 and Statement 2 are based on the homogeneous estimate for the activation barrier height. These properties can be violated. But as far as these Statements are based on very strong inequalities one can accept their validity. The influence of nonstationary effects will be analyzed in a separate publication where the complete theory has been constructed. Now one can analyze the profile of intensity of the droplets formation around the already existing droplet. This nucleation rate profile is a rather sharp function which has a step-like behavior. To show this we shall introduce two characteristic values of $`\beta `$ ($`\beta _{st}`$ and $`\beta _{fin}`$) by relations $$f(\beta _{st})=\sqrt{\frac{\pi }{2}}\sqrt{\frac{v_v}{v_l}}\frac{\mathrm{exp}(1/2)}{\mathrm{\Gamma }}$$ $$f(\beta _{fin})=\sqrt{\frac{\pi }{2}}\sqrt{\frac{v_v}{v_l}}\frac{\mathrm{exp}(1/2)}{\mathrm{\Gamma }}$$ In the region $`\beta >\beta _{st}`$ the rate of nucleation practically coincides with the unperturbed value $`I_s(\zeta _0)`$. In the region $`\zeta <\zeta _{fin}`$ the rate of nucleation is negligible in comparison with the unperturbed value, i.e. $`I_s(\zeta (r))I_s(\zeta _0)`$. In some moment $`t`$ the values $`\beta _{st}`$, $`\beta _{fin}`$ initiate the space distances $`r_{st}`$, $`r_{fin}`$ by the following expressions $$r_{st}=\beta _{st}\sqrt{4Dt}$$ $$r_{fin}=\beta _{fin}\sqrt{4Dt}$$ For $`\sigma 1`$ one can come to $$f(\beta _{st})1$$ $$f(\beta _{fin})1$$ and $$\beta _{st}1$$ $$\beta _{fin}1$$ Then one can use the asymptote (8) and see that $$\frac{|\beta _{st}\beta _{fin}|}{\beta _{st}+\beta _{fin}}=\frac{1}{4\beta _{st,fin}}1$$ $$\frac{|r_{st}r_{fin}|}{r_{st}+r_{fin}}=\frac{1}{4\beta _{st,fin}}1$$ The real picture of nucleation is going in the time scale. In a fixed space point $`r`$ one can introduce two characteristic times $`t_{st}`$ and $`t_{fin}`$ by expressions $$t_{st}=\frac{r^2}{4\beta _{st}^2D}$$ $$t_{fin}=\frac{r^2}{4\beta _{fin}^2D}$$ Before $`t_{st}`$ one can not observe any essential deviation of the nucleation rate from the unperturbed value. After $`t_{fin}`$ the rate of nucleation is very small. One can get for the relative deviation $$\delta \frac{t_{fin}t_{st}}{t_{st,fin}}$$ the following expression $$\delta \frac{1}{\beta _{st,fin}}$$ So, the relative deviation is small. Even in the situation of small $`\sigma `$ one can show with the help of the asymptote (7) that $`\delta `$ is rather small. The step-like behavior<sup>15</sup><sup>15</sup>15This step can take place even for $`\beta >\beta _{st}`$ when $`\sigma 1`$. of intensity profile allows to introduce some characteristic parameter $`\beta _{eff}`$ and consider the region<sup>16</sup><sup>16</sup>16At $`\sigma 1`$ the value of $`\beta _{eff}`$ can be greater tan $`\beta st`$ and $`\beta _{fin}`$. $$\beta <\beta _{eff}$$ as the exhausted region where there is no nucleation more and in the region $$\beta >\beta _{eff}$$ the rate of nucleation is unperturbed<sup>17</sup><sup>17</sup>17In all cases $`\beta _{eff}>\beta _{fin}`$.. One has to choose $`\beta _{eff}`$ carefully. The problem is the possibility of existence of a long tail of a density profile. To grasp the situation of a small values of $`\sigma `$ one has to introduce $`\beta _{eff}`$ in an integral manner. One can introduce the excess of nucleation rate $`\mathrm{\Delta }I_s`$ by the following formula $$\mathrm{\Delta }I_s=I_s_0^{\mathrm{}}(1\mathrm{exp}(\frac{\mathrm{\Gamma }(\zeta _0\zeta (r))}{\zeta _0}))4\pi r^2𝑑r$$ when $`I_s`$ is the unperturbed rate of nucleation. On the base of the last expression one can get the excess of $`N`$ due to the existence of the solitary profile. This value will be marked by $`\mathrm{\Delta }N_{sol}`$ and can be found as $$\mathrm{\Delta }N_{sol}=I_s_0^t_0^{\mathrm{}}(1\mathrm{exp}(\frac{\mathrm{\Gamma }(\zeta _0\zeta (r))}{\zeta _0}))4\pi r^2𝑑r𝑑t^{}$$ Having used equation (6) one can come to $$\mathrm{\Delta }I_s=4\pi (4Dt)^{3/2}I_s_0^{\mathrm{}}(1\mathrm{exp}(\mathrm{\Gamma }\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}f(\beta )))\beta ^2𝑑\beta $$ Parameter $`\mathrm{\Gamma }\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}`$ has a constant value. The value $`\mathrm{\Delta }N_{sol}`$ can be now found as $$\mathrm{\Delta }N_{sol}=4\pi (4Dt)^{3/2}I_s_0^t_0^{\mathrm{}}(1\mathrm{exp}(\mathrm{\Gamma }\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}f(\beta )))\beta ^2𝑑\beta 𝑑t^{}$$ The step-like approximation of the nucleation profile will lead to $$\mathrm{\Delta }I_s^0(\beta _{eff})=4\pi (4Dt)^{3/2}I_s_0^{\beta _{eff}}x^2𝑑x$$ The value $`\beta _{eff}`$ has to be determined from $$\mathrm{\Delta }I_s^0(\beta _{eff})=\mathrm{\Delta }I_s$$ Certainly, the value of $`\beta _{eff}`$ depends of $`\mathrm{\Gamma }\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}`$. The value of $`\beta _{eff}`$ initiates $$r_{eff}=2\beta _{eff}D^{1/2}t^{1/2}$$ Inside the volume $$V_{eff}=\frac{4}{3}\pi r_{eff}^3$$ there is no nucleation and outside this volume the rate of nucleation is unperturbed. Then one can imagine that around every droplet there exists the exhausted region (ER) where there is no nucleation and the unexhausted region (UR) where the nucleation remains unperturbed. The whole space now is divided into two regions. The volume $`V_{eff}`$ grows in time in the following way $$V_{eff}=\frac{32}{3}\pi \beta _{eff}^3D^{3/2}t^{3/2}$$ In a free molecular regime $`V_{eff}`$ will grow even faster. For $`b_{eff}`$ one can get the simple expression $$\beta _{eff}^3=3_0^{\mathrm{}}(1\mathrm{exp}(\mathrm{\Gamma }\sqrt{\frac{2}{\pi }}\sqrt{\frac{v_l}{v_v}}f(\beta )))\beta ^2𝑑\beta $$ or $$\beta _{eff}^3=3_0^{\mathrm{}}(1\mathrm{exp}(\sigma ^{1/2}\sqrt{\frac{2}{\pi }}f(\beta )))\beta ^2𝑑\beta $$ For $`\mathrm{\Delta }N_{sol}`$ one can obtain $$\mathrm{\Delta }N_{sol}=I_s(\zeta _0)_0^t𝑑t^{}V_{eff}=_0^t𝑑t^{}\frac{4}{3}\pi r_{eff}^3$$ One can easy integrate the last expression and get $$\mathrm{\Delta }N_{sol}=I_s(\zeta _0)\frac{64}{15}\pi \beta _{eff}^3D^{3/2}t^{5/2}$$ One can see that $`\mathrm{\Delta }N_{sol}`$ is growing in time rather rapidly. Namely this property illustrates the feature of the ”avalanche consumption during the first order phase transition” in application to the heterogeneous nucleation. For those situations where $`\sigma 1`$ one can get $$\beta _{eff}\beta _{st}\beta _{fin}$$ and $`\beta _{eff}`$ satisfies the simple equation $$\mathrm{exp}(\beta _{eff}^2)=\beta _{eff}^3\sqrt{\frac{v_v}{v_l}}\sqrt{\frac{\pi }{2}}\frac{1}{\mathrm{\Gamma }}$$ The last equation can be easily solved by iterations as far as $`\beta _{eff}1`$ and $`\mathrm{exp}(\beta ^2)`$ is the most sharp function of the supersaturation. Earlier when the principle of the separate growth was discussed then the reason of the absence of interaction between droplets was the low probability to appear too close one to another only due to the smallness of such space volume. Now one can see that the growing ER helps to exclude the interference. Really, the essential deviation of the supersaturation from the ideal one can be seen in the region $`r<R_dl`$. It means that the distance between the droplets with an interference must have the order $`2R_dl`$ Then the time distance between the moments of formation of these droplets must be shorter than $$\mathrm{\Delta }t_{init}(\frac{R_dl}{\beta _{eff}D^{1/2}})^2$$ This time interval is many times shorter than the duration of the nucleation period. Every droplet forms rather rapidly after formation the ER of such a size which guarantees that the rate of growth of the given droplet can not be perturbed by a vapor consumption by the other droplets. ## 2 Kinetic models of the global evolution Now one can construct the picture of nucleation in a whole system. The main problem is to take into account the interference of the density profiles. The interference through the rate of growth is absent, but there is a simple overlapping of profiles which leads to the deviation of the total nucleation rate over the volume from those calculated with account of the additive excesses around every droplet. The overlapping of ER (even when this approximate formalism is used) is very complex and can not be directly taken into account in precise manner. Instead of writing the long expressions which can not be explicitly calculated one can act in another manner: some simple approximate models for kinetics of the nucleation process will be formulated. From these models it will be seen that they estimate the nucleation characteristics from below and from above and lead practically to the similar results. So, it will be shown that the complex details of the ER overlapping has no strong influence on the real characteristics of the phase transition. At first one can consider the common feature of all models concerning the exhaustion of free heterogeneous centers. The rate of nucleation $`I`$ depends on time $`t`$ and on spatial point $`r`$ (the last behavior is the most complex). So, it is reasonable to consider the mean (over the space) value of $`I`$ denoting it by $`<I>`$. For $`<I>`$ one can write the following expression $$<I>=\frac{W_{free}}{W_{tot}}\frac{\eta }{\eta _{tot}}I_0$$ (9) where $`I_0`$ is the unperturbed rate of nucleation. Here $`W_{free}`$ is the volume of a region where the rate of nucleation is unperturbed, i.e. the UR of the whole system. The value $`W_{tot}`$ is the total volume of the system (it equals to unity and is written only to clarify the consideration). Then as far as $$N=_0^t<I>(t^{})𝑑t^{}$$ one can get $$\eta =\eta _{tot}_0^t<I>(t^{})𝑑t^{}$$ In the differential form the last relation can be written as $$\frac{d\eta }{dt}=<I>$$ and with the help of (9) it can be rewritten as $$\frac{d\eta }{dt}=\frac{\eta }{\eta _{tot}}\frac{W_{free}}{W_{tot}}I_0$$ After the integration of the last expression one can come to $$\eta =\eta _{tot}\mathrm{exp}(_0^t\frac{W_{free}(t^{})}{W_{tot}}\frac{I_0}{\eta _{tot}}𝑑t^{})$$ One has to note that heterogeneous centers aren’t distributed homogeneously with respect to the ER (or UR). Only free heterogeneous centers are distributed homogeneously with respect to ER. This fact has to be also taken into account. The problem is to determine the value of $`W_{free}`$. In different models it will be determined in different forms. The model without overlapping. One can write $$W_{free}=W_{tot}W_{exh}$$ where $`W_{exh}`$ is the volume where there is no further formation of the droplets. Very approximately one can present it as the sum of ER around all appeared droplets $$W_{exh}\underset{i}{}V_{eff}$$ (the sum is taken over all already formed droplets). Certainly, the last approximation is rigorous only when there is no overlapping of the ER around different droplets. Having used an expression for $`V_{eff}`$ one can come to $$W_{exh}=_0^t𝑑t^{}<I>(t^{})\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}$$ (10) Having used an expression for $`<I>`$ one can come to the closed system of nucleation kinetics equations $$W_{free}=W_{tot}_0^t𝑑t^{}\frac{\eta }{\eta _{tot}}\frac{W_{free}}{W_{tot}}I_0\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}$$ $$\eta =\eta _{tot}\mathrm{exp}(_0^t\frac{W_{free}(t^{})}{W_{tot}}\frac{I_0}{\eta _{tot}}𝑑t^{})$$ (11) In the quasihomogeneous limit (when there is no essential exhaustion of centers) this system can be reduced to $$W_{free}=W_{tot}_0^t𝑑t^{}\frac{W_{free}}{W_{tot}}I_0\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}$$ which can be written after the evident renormalization $`tat,t^{}at^{}`$ where $`a=(I_0\frac{32}{3}\pi \beta _{eff}^3D^{3/2})^{2/5}`$ in the universal form $$W_{free}=1_0^t𝑑t^{}W_{free}(tt^{})^{3/2}$$ One has to note that in the general case the system of nucleation equations can be solved with the help of methods developed in . At first one can solve the quasihomogeneous equation (it the Volterra equation with a rather simple kernel which allows to apply the Laplace transformation to solve it) and then on the base of quasihomogeneous equation one can find the final rather precise expression using the second equation as a formula for $`\eta `$. Another variant is to solve numerically the universal equation for $`W_{freehom}`$: $$W_{freehom}=1_0^t𝑑t^{}W_{freehom}(tt^{})^{3/2}$$ Then one has the universal function $`W_{freehom}`$. Then one can find $`\eta `$ as $$\eta =\eta _{tot}\mathrm{exp}(a^1_0^t\frac{W_{freehom}(t^{})}{W_{tot}}\frac{I_0}{\eta _{tot}}𝑑t^{})$$ The last expression leads to the formula for $`<I>`$ as: $$<I>=\frac{W_{freehom}}{W_{tot}}\mathrm{exp}(a^1_0^t\frac{W_{freehom}(t^{})}{W_{tot}}\frac{I_0}{\eta _{tot}}𝑑t^{}I_0)$$ The justification of such approach is analogous to . The physical reason is very simple: when there is no exhaustion of heterogeneous centers then the solution is found precisely, when there is an essential exhaustion of centers there is no need to know $`W_{free}`$ with a high precision because the converging force of the operator of the right hand of the second equation in the nucleation equations system is extremely high. Now we shall take into account the overlapping. It can be done rather approximately. The model with chaotic overlapping The matter of discussion is the correct expression for $`W_{free}`$ which can not be got absolutely precisely. Now the more reasonable expression for $`W_{free}`$ will be presented. Certainly, this will lead to a more complicate equation which will be more difficult to solve. One can use the differential approach to write the expression for $`W_{free}`$. Having written the evident relation $$\frac{dW_{free}}{dt}=\frac{dW_{exh}}{dt}$$ one has to invent an approximation for $`dW_{exh}/dt`$. Here the following approximation $$\frac{dW_{exh}}{dt}\frac{d_iV_{eff}}{dt}\frac{W_{free}}{W_{tot}}$$ (the sum is taken over all droplets) will be used. It corresponds to the following approach: The probability of the absence of overlapping of the new parts of ER around the given droplet with other ER is proportional to the free volume of the system. The last supposition seems to be rather reasonable. The value $`d_iV_{eff}/dt`$ can be rewritten as $$\frac{d_iV_{eff}}{dt}=\underset{i}{}\frac{dV_{eff}}{dt}$$ The last value can be easily expressed through $`<I>`$ as $$\underset{i}{}\frac{dV_{eff}}{dt}=\frac{3}{2}_0^t𝑑t^{}<I>(t^{})\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}$$ (12) due to (10). Then $$\frac{dW_{exh}}{dt}=\frac{3}{2}_0^t<I>(t^{})\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}𝑑t^{}W_{free}(t)$$ and $$\frac{dW_{free}}{dt}=\frac{3}{2}_0^t<I>(t^{})\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}𝑑t^{}W_{free}(t)$$ Having used an expression for $`<I>`$ one can come to $$\frac{dW_{free}}{dt}=\frac{3}{2}_0^t\frac{W_{free}}{W_{tot}}I_0\frac{\eta }{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}𝑑t^{}W_{free}(t)$$ Together with (11) the last equation forms the closed system of the nucleation equations in the second model. The previous equation can be integrated which gives $$\mathrm{ln}W_{free}=_0^t\frac{W_{free}(t^{})}{W_{tot}}I_0\frac{\eta (t^{})}{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}𝑑t^{}+const$$ Due to the initial conditions the $`const`$ is equal to zero. Having introduced the function $`F=\mathrm{ln}W_{free}`$ one can get for $`F`$ the following system of equations $$F(t)=_0^t\mathrm{exp}(F(t^{}))I_0\frac{\eta (t^{})}{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}𝑑t^{}$$ $$\eta =\eta _{tot}\mathrm{exp}(_0^t\mathrm{exp}(F(t^{}))\frac{I_0}{\eta _{tot}}𝑑t^{})$$ One can see that the system of condensation equations is identical to the system of nucleation equations in AA. It was completely analyzed in . Certainly, the parameters in the system will be another. The last system can be rewritten after the evident renormalization as $$F(t)=_0^t\mathrm{exp}(F(t^{}))(tt^{})^{3/2}\theta (t^{})𝑑t^{}\widehat{F}(F,\theta )$$ $$\theta (t)=exp[A_0^t\mathrm{exp}(F(t^{}))𝑑t^{}]\widehat{\theta }(F)$$ where $`\theta (t)=\eta (t)/\eta _{tot}`$ and $`A`$ is some known parameter. This system can be solved by iterations defined as $$F_{i+1}=\widehat{F}(F_i,\theta _i)$$ $$\theta _{i+1}=\widehat{\theta }(F_i)$$ with $`F_0=0`$, $`\theta _0=1`$. For $`F_i`$, $`\theta _i`$ one can get the chains of inequalities $$F_0<F_2..<F_{2i}<\mathrm{}<F<\mathrm{}<F_{2i+1}<\mathrm{}<F_3<F_1$$ $$\theta _1<\theta _3<\mathrm{}<\theta _{2i+1}<\mathrm{}<\theta <\mathrm{}<\theta _{2i}<\mathrm{}<\theta _2<\theta _0$$ Then one can estimate errors in $`F_i`$, $`\theta _i`$. One can use also another methods analogous to those described in . The similarity of the the condensation equations in AA and in the second model is extremely important for the transition towards collective character of vapor consumption which is analyzed in . The physical ground of the considered model is the chaotic overlapping of ER. Namely, the chaotic overlapping lies in the base of approximation used here. But due to the spherical form of every ER the overlapping isn’t absolutely chaotic. What can be done in such a situation? In the next model we shall show that concrete type of overlapping isn’t very important. To finish with the second model we shall show the same method of its solution as that of the first model. One can also formulate the quasihomogeneous equation as the following one $$F_{hom}(t)=_0^t\mathrm{exp}(F_{hom}(t^{}))I_0\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}𝑑t^{}$$ Then $`\eta `$ can be approximately found as $$\eta =\eta _{tot}\mathrm{exp}(_0^t\mathrm{exp}(F_{hom}(t^{}))\frac{I_0}{\eta _{tot}}dt^{}$$ The quasihomogeneous equation can be renormalized. After renormalization $`zat`$, $`t^{}at^{}`$ where $$a=(I_0\frac{32}{3}\pi \beta _{eff}^3D^{3/2})^{2/5}$$ one can transform the quasihomogeneous equation into the universal form $$\mathrm{ln}W_{freehom}(t)=_0^tW_{freehom}(t^{})(tt^{})^{3/2}𝑑t^{}$$ The model with the formation of droplets inside the ER The third model will show that the role of overlapping isn’t so essential as it can be imagined from the first point of view. Suppose that the new droplets can appear also in the ER of the already existed droplets. Then instead of (12) one has to use $$\underset{i}{}\frac{dV_{eff}}{dt}=\frac{3}{2}_0^t𝑑t^{}I_0\frac{\eta (t^{})}{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}$$ Then $$\frac{dW_{exh}}{dt}=\frac{3}{2}_0^t\frac{\eta (t^{})}{\eta _{tot}}I_0\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}𝑑t^{}W_{free}(t)$$ and $$\frac{dW_{free}}{dt}=\frac{3}{2}_0^tI_0\frac{\eta (t^{})}{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{1/2}𝑑t^{}W_{free}(t)$$ Together with (11) the last equation forms the closed system of the nucleation equations in the third model. The substance balance equation of the system can be integrated which gives $$\mathrm{ln}W_{free}=_0^tI_0\frac{\eta (t^{})}{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}𝑑t^{}+const$$ Due to the initial conditions the $`const`$ is equal to zero. Having introduced the function $`F=\mathrm{ln}W_{free}`$ one can get for $`F`$ the following system of equations $$F(t)=_0^tI_0\frac{\eta (t^{})}{\eta _{tot}}\frac{32}{3}\pi \beta _{eff}^3D^{3/2}(tt^{})^{3/2}𝑑t^{}$$ $$\eta =\eta _{tot}\mathrm{exp}(_0^t\mathrm{exp}(F(t^{}))\frac{I_0}{\eta _{tot}}𝑑t^{})$$ This system corresponds to the first iteration in the iteration solution of the second model by the method of iterations of a type ”a” described in . It is also described here. But for $`\eta `$ the whole set of iterations has been taken (see details in ). One can slightly modify the model and suppose that in the expression for $`\eta `$ one can use the same approximation for $`<I>`$ as in equation for $`W_{free}`$. Then the system of equation will precisely correspond to the first iteration in the iteration solution in . Actually one needn’t to analyze these models in details following but has only to note that all these solutions are practically similar. One can see the solution of the third model (both variants) after the same renormalizations as in the second model. Now one has to explain why the third model is rather accurate. One can do it only with the help of results obtained in . In it was noted that when the power of the kernel $`(tt^{})`$ is rather big then the solution of the quasihomogeneous equation weakly depends on concrete value of a power. Another important notation shows that when the power of $`(tt^{})`$ is extremely high then the ER of the given (first) droplet formed inside the ER of another (second) droplet practically can not go outside the ER of that (second) droplet. Then the third model is absolute adequate in this situation. The same feature can be seen directly from the iteration procedure results. Having combined these two notations one can see that on one hand the second model is close to the situation with great power of the kernel and in the situation with the big kernel the third model is suitable. Now it is possible to explain why the overlapping isn’t so important as it can be imagined. Really, as far as the power of kernel is big and one can observe the avalanche consumption one can see the following qualitative picture * Practically during all period of nucleation the total ER is small and there is no problem of overlapping. * At the end of the nucleation period the total ER is going to occupy the essential part of the volume and a few moments late it occupies all the volume of the system. This process is very rapid. It stops the nucleation. This picture shows that there is no strong influence of the overlapping on the nucleation process (except the cut-off). As the result the nucleation description is completed. One can use both the second and the third model to get the nucleation description. How to solve these equations is also described here. The method of universal solution The main idea of the theory presented in was to consider the quasihomogeneous equation, to get the universal solution and then on the base of this solution to calculate the number of free heterogeneous centers. As the result one can get an expression for $`<I>`$ and can calculate the total number of droplets appeared on the nucleation process. Here one can follow the same idea. But one can make this idea more deep. One needn’t to formulate the universal quasihomogeneous equation. Instead of equation one can formulate the universal model. The model will be the following * -the rate of nucleation $`<I>`$ can be found from $$<I>=I_0\frac{W_{free}}{W_{tot}}\frac{\eta }{\eta _{tot}}$$ * \- with intensity $`I_0`$ the droplet appears in the arbitrary point of the system which belongs to the free volume. * -the value $`W_{free}`$ can be found by exclusion of all ER around the already existing droplets. * \- if the point is occupied by the ER of any droplet then the droplet can not be formed. * \- the size $`r_{eff}`$ of ER grows in time according to $$r_{eff}=2\beta _{eff}D^{1/2}t^{1/2}$$ * -the initial condition is the absence of droplets and random distribution of centers. With the proper renormalization of time $`t`$ and size $`r`$ one can cancel all coefficients. Then this process will be the universal one and as the result $`W_{free}`$ is the universal function of time. Then one can directly apply (11) and get the number of free heterogeneous centers (after the proper renormalization)<sup>18</sup><sup>18</sup>18This is quasihomogeneous approach method. The modification for the dynamic conditions requires to put instead of $`I_0`$ the value $`I_0exp(ct)`$ with some parameter $`c`$ determined by external conditions and to change the lower limit $`0`$ of integration to $`\mathrm{}`$ . The main constructions of the theory will be absolutely the same but the forms of characteristic curves changes radically. ## 3 Numerical results Numerical simulation in the nucleation kinetics plays at least two important roles. The first role is the standard comparison with the approximate models to observe their quality. The second one is more specific and it is concerned with some universal dependencies in the nucleation kinetics. In the AA to the nucleation kinetics it was shown that the adequate approach can be presented on the base of the quasihomogeneous solution . Despite the dynamic conditions considered in one can observe this property in the situation of the metastable phase decay also. Recall the reasons for such approach. The formal reason is the careful analysis of the iteration procedure proposed in . Really, the final result for the total number of droplets appeared in the nucleation process is given by the second iteration (see iterations of type ”a” in ) for the relative number of free heterogeneous centers. This iteration is based only on the first iteration for the supersaturation. The value of supersaturation there is calculated without any account of the heterogeneous centers exhaustion. So, one can see that the final result can be obtained on the base of the supersaturation in the quasihomogeneous approximation. This approximation can be more sophisticated than the first iteration. Namely this approach was used in where the precise quasihomogeneous universal solution was chosen as base for the final results. The physical reason of such behavior is rather simple. The main role in the vapor consumption belongs to the droplets of relatively large sizes. We have already marked this fact. But moreover, due to the avalanche character of the vapor consumption the main role in vapor consumption is played by the relatively large droplets which are formed in the first moments of time of the nucleation period. When the effect of the centers exhaustion is essential already in the first moments of nucleation period<sup>19</sup><sup>19</sup>19More precisely one can define these ”first moments of time” as $`2/5`$ of the nucleation period duration (under the free molecular regime it is $`1/4`$ of the nucleation period duration). The reason for concrete value can come from the iteration procedure. then at the end of nucleation period already all centers will be exhausted. Then the result is evident - all centers are the centers of droplets. Due to the high force of convergence in this situation this result can be gotten under the rather arbitrary behavior of supersaturation (even including the quasihomogeneous case). In the opposite case the exhaustion of heterogeneous centers during the first moments of the nucleation period isn’t essential and one can get the quasihomogeneous behavior of supersaturation. This effect will be called as the ”approximate separation of the heterogeneous and homogeneous problems”. It is based only on the avalanche consumption of the metastable phase. So, there are no objections to see this effect also in the situation with the density profiles. Then it is extremely important to get solution in the quasihomogeneous situation and clarify whether it can be presented in the universal form. The universal form of the quasihomogeneous solution can be easily seen in the situation of profiles also. In the AA there was no specific space scale because the consumption took place homogeneously from all space points of the system. Here in the situation of profiles there is the elementary space scale and one can choose the space scale as to have that the linear size of ER around the droplet is growing<sup>20</sup><sup>20</sup>20Certainly, the power has to be conserved. as $`t^{1/2}`$. The time scale has to be chosen as to get that in the free volume equal to the total volume of the system one can see the appearance of one droplet in the unit of time. As long as the functional dependencies of the nucleation rate and the radius of ER are not identical one can do such a renormalization without any problems. So, we see that the pseudohomogeneous case allows the universal description. The process of the heterogeneous centers exhaustion destroys this universality and one has to act as in<sup>21</sup><sup>21</sup>21Here we use slightly more simple way. . The total number of droplets has to be approximately calculated as $$N_{total}=\eta _{tot}[1\mathrm{exp}(\frac{N_{hom}}{\eta _{tot}})]$$ (13) where $`N_{hom}`$ is the number of the droplets appeared in the quasihomogeneous situation (under the same parameters). This formula can be also used for all approximate models described earlier. For the numerical simulation it was convenient to consider the cube box with a size $`10`$. The rate of the ER growth is chosen as $$\frac{dR}{dt}=100t^{1/2}$$ where $`R`$ is the radius of ER. The rate of nucleation is chosen as to have one attempt of a new droplet formation in the system during $`dt=0.002`$. This attempt is governed by the random procedure. It may lead to the point in one of the ER and then no droplet will be formed. In the opposite situation when the point is indicated out of all ER of the already existing droplets there will be a new formation of a droplet. One has to stress that the random procedure has one specific negative feature. In the standard numerical procedures the next random coordinate is calculated on the base of the previous ones. So, if the current coordinate lies near the center of the already existing ER then the next coordinate will be also near the center of another ER. These correlations lead to the necessity to consider the large system. In the system under consideration the number of droplets appeared in the quasihomogeneous situation will be near $`500`$. Nevertheless the square mean fluctuation will be about $`20`$. The careful consideration can show that the error induced by the substitution of the periodic boundary conditions by the zero ones has the same power as the mean square error. It can be directly seen by the numerical simulation. It is explained by the evident notation that the characteristic overlapping of profiles is about the mean profile size. We shall call this feature as the property of ”moderate overlapping”. This fact can be proven analytically. One can also analytically prove that the effect of fluctuations is really caused by the mentioned phenomena. The mean value of the total droplets number is equal to $`504.8`$ (under the zero boundary conditions). This value has to be put into the previous formula. The avalanche character of the vapor consumption is illustrated by fig. 2-4. Three different moments of time $`t=0.5`$, $`t=1`$ and $`t=1.5`$ are chosen as characteristic values. The cross section is drawn. The dashed regions correspond to the ER of the already existing droplets. The black regions correspond to the overlapping<sup>22</sup><sup>22</sup>22Only when the distance between centers is approximately the odd one. of ER. The number of heterogeneous centers in such a system can be arbitrary. Certainly the effects of the centers exhaustion will be important when the number of centers is small (in comparison with $`N_{hom}=504.8`$). The pictures for $`\eta _{tot}=50`$ are drawn in fig. 5-7 for $`t=1.5`$, $`t=3`$, $`t=6`$. One can see that the number of ER is smaller than in the quasihomogeneous case. The size of ER when the free volume is going to be exhausted is larger than the size of ER in the quasihomogeneous case. The time necessary to cover all volume by ER is greater than in the quasihomogeneous case. It doesn’t mean that the duration of the nucleation period will be longer (simply all centers will be exhausted and this means the end of nucleation). Moreover, the duration of the nucleation period in the situation of the relatively small number of heterogeneous centers will be shorter than in the quasihomogeneous case. One can also see that the avalanche character of the vapor consumption here will be more smooth than in the quasihomogeneous case. Certainly, in the quasihomogeneous case the appearance of the new ER helps to consume the vapor phase in the avalanche manner. But in the situation with the small number of centers there is no need to consider the process carefully because the exhaustion of centers leads to the evident result of condensation \- the number of the droplets equals to the number of centers. Now it is evident that the main object of our interest will be the quasihomogeneous case. The relative rate of nucleation in this case is shown in fig. 8. Here the rate of nucleation is averaged over $`100dt0.2`$ and over $`16`$ attempts. So, the rate of nucleation is rather smooth function. The relative rate of nucleation is compared in fig.8 with the mentioned models. Certainly, the rate of nucleation defines the spectrum of sizes when we take as the size of embryo those characteristic which has the rate of growth independent on the value of size. For the diffusion regime this characteristic is the number of molecules in the power $`2/3`$. One can see in fig. 8 three different curves and some solitary points. The solitary points correspond to the numerical simulation of the quasihomogeneous case and three curves correspond to three models in the quasihomogeneous case. The shortest spectrum is in the first model. The careful look shows that this line is doubled. This occurs because there the ideal variant of the first model is also drawn. This ideal variant corresponds to $`W_{free}/W_{total}1`$ in the subintegral function. The coincidence means that the main role in the first model is played by the relatively big droplets which were formed at $`W_{free}=W_{total}`$. The longest spectrum corresponds to the second model. This curve is very close to the intermediate curve which corresponds to the third model. The approximate coincidence of the second and the third models shows that both of them are valid and the role of the relatively big droplets here is the main one. One also see that the first model isn’t too far from the real solution. This allows to present the rigorous estimates for the nucleation rate. Certainly the first model is the estimate of the real process from below. It gives the number of the droplets about $`20`$ percents less than the numerical simulation. An estimate for the nucleation rate from above can be gotten by the following way. From the first model it follows that until $`t=0.52`$ (practically this case is drawn in fig. 2) the rate of nucleation is near the ideal value and the deviation is less than $`15`$ percents. So, one can say that the period $`0<t<0.52`$ corresponds to the absence of overlapping (the first model is the estimate from above). Then one can consider the process where the total volume is exhausted only by the ER of the droplets appeared at $`0<t<0.52`$ in a random manner. The distribution of the centers of ER of such droplets is evidently random. This model certainly gives the estimate from above for the nucleation process. The total number of the droplets is only $`15`$ percents greater than the result of the numerical simulation. As a conclusion one can state that two suitable estimates from below and above are obtained. The proximity of the last estimate to the real solution justifies the supposition that the main role in vapor consumption belongs to the droplets of the relatively big sizes appeared in the system practically free from the ER. This supposition can be also justified in the analytical manner. One can see that the second and the third models are rather close to the real solution but don’t coincide with it. There are at least two reasons of such deviation. The first one is the presence of the strong correlations in the real system - if two ER overlap in some moment of time then the power of overlapping can only grow in time. It hasn’t a random character as it is stated in the second and third models. This effect can be taken into account in a rather simple manner. It is sufficient to consider two spheres and calculate the power of overlapping as the function of the distance and time (it is the simple geometrical problem). Unfortunately the answer can be written only in a very complicate form. If we have two ER with radii $`R_1`$ and $`R_2`$ correspondingly at a distance $`l`$ between their centers and $`l>max(R_1,R_2)`$ then the volume of overlapping is $$V_{over}=\frac{2\pi R_1^3}{3}(12cos\phi _1+cos^3\phi _1)+\frac{2\pi R_2^3}{3}(12cos\phi _2+cos^3\phi _2)$$ where $$cos\phi _1=\frac{R_2^2+R_1^2+l^2}{2R_1l}$$ $$cos\phi _2=\frac{R_1^2+R_2^2+l^2}{2R_2l}$$ Certainly, this result can not lead to the compact form of the balance equation. It will be difficult to solve it analytically. The second reason is the ”moderate overlapping” problem. This property means that actually there is an interaction through overlapping in ensemble of several droplets. Earlier this property was extracted in a case of a special effective length of ER. Now we see that this property has a rather general sense. The way to solve this problem proposed in is very complicate and leads to some uncertain relations. How one can overcome all these problems? Actually one has no need to do it. The simple numerical simulation takes into account all these effects and gives the universal solution. Really we need only one number - the total number of appeared droplets. It is known. Then one can forget about all mentioned difficulties. Now one can analyze the heterogeneous case explicitly. The suitable approximation is given by (13). One has to substitute instead of $`N_{hom}`$ the number of droplets given by the corresponding model. The relative error of approximation (13) is drawn in fig. 9 for the first model, in fig. 10 for the second model, in fig. 11 for the third model. It is rather small for all models. For the third model it is practically negligible. It is clear because the third model is based on the free volume approximation. Then the source of error can be found only in exhaustion of centers without any influence on the growing of ER. Without this influence the power of exhaustion can not essentially act on the process of nucleation. One can fulfill the same analysis for the numerical simulation. In fig. 12 the relative error of (13) for numerical simulation is drawn. Here in (13) the value $`N_{hom}=504.8`$ from the numerical simulation is used. The result is compared with the computer simulation of heterogeneous condensation. This simulation is rather simple. One can take the procedure for the quasihomogeneous case but put the center of the new droplet with probability $`\eta /\eta _{tot}`$. Every time when this point will be out of ER we shall reduce $`\eta `$ as $`\eta \eta 1`$. One can see that the relative error is very small. We don’t use the averaging over (this is the reason why there is no smooth curve) many attempts in order to see that the error of (13) has the scale of the mean square error of numerical simulation<sup>23</sup><sup>23</sup>23For the system with $`500`$ droplets.. So, there is no need to use the more sophisticated approach. The solution of the problem now is completed. The generalization for the conditions of dynamic type is absolutely analogous to . The convergence due to avalanche consumption is more weak and one has to use instead of approximation (13) more sophisticated procedure described in . Certainly, the universal constants have to be calculated by the numerical simulation with profiles. The generalization for the arbitrary regime of the droplets growth can be done absolutely analogous to . This generalization is based on the similarity of the functional forms obtained here and in the AA. This similarity lies in the base of universality formulated in , . One can see that the theory of condensation with profiles taken into account presents the picture which is absolutely different from the AA. Nevertheless in many situations the result of experiment coincides with the result of the AA. One has to explain this coincidence. It is rather formal one. Really, in any experiment it is more convenient to have a small system and to get many droplets. The rate of nucleation has to be taken as a rather high one. So, the supersaturation is relatively high and parameter $`\sigma ^1`$ isn’t a real small parameter of a theory<sup>24</sup><sup>24</sup>24 It isn’t necessary for consideration here, but it has to be small due to the possibility of thermodynamic description of the critical embryo.. Then as it is shown in the AA gives the write qualitative result despite the wrong base of consideration. The reason lies in the fact that at small $`\sigma `$ the main quantity of a substance is lying in the tail of profile. The tail of profile is rather thin and can be be taken into account by the AA. In the correction term for the AA at small $`\sigma `$ is also presented. The last important feature to mention is the movement of the embryos boundaries. This problem is widely discussed in the determination of the rate of regular growth for the supercritical embryos. In different systems the effect of influence of the boundary movement on the rate of growth is different. We have to note that in theory presented here the rate of the embryos growth is an external value which is assumed to be known<sup>25</sup><sup>25</sup>25It is really known practically for all systems.. Another problem is the adequate account of the effect of boundary movement in construction of the ER. One can see that in the first part of publication it is already shown that the effect of the boundary movement is small. Here we shall present the abstract arguments for such conclusion. To use the thermodynamic approach the initial power of the mother phase metastability has to be relatively small. Together with the Maxwell’s rule it leads to the following final result of the phase transition: ”Only relatively small part of the system volume is occupied by the new phase.” It isn’t in contradiction with the property that all volume is occupied by ER. So, the final state of the system is the practically saturated mother phase in practically all volume of the system and the small volume (distributed over all system) occupied by a new phase. As the result one can see that the process of a substance consumption (extraction) leads to the saturation in a volume relatively large in comparison with a volume of the new phase embryo. The mother phase can not be undersaturated (then there will be the disappearance of embryos). So, as long as even the saturated phase has to be spread on large distances $`l_0`$ one can state that the growth of embryo produces the perturbation over the large distances $`l>l_0`$. To have an interruption (relative interruption in comparison with an ideal nucleation rate) of a new phase formation one needs a very small reduction of the power of metastability<sup>26</sup><sup>26</sup>26Relative reduction has to be $`\mathrm{\Gamma }^1`$ where $`\mathrm{\Gamma }1`$ has the scale of the number of molecules in the critical embryo.. So, this reduction is attained at the distances $`l^{}>l`$ which are very large in comparison with the embryos linear size. Then one can use the point source approximation as it was done in the first part and forget about the boundaries movement. The negligible character of the boundary movement is proven now for all possible systems. The heat extraction and account of all other intensive parameters of description can be done analogously to . Fig. 1 The form of $`f(\beta )`$. Fig. 2 The cross section of the system at $`t=0.5`$ for the quasihomogeneous situation. Fig. 3 The cross section of the system at $`t=1`$ for the quasihomogeneous situation. Fig. 4 The cross section of the system at $`t=1.5`$ for the quasihomogeneous system. Fig. 5 The cross section of the system at $`t=1.5`$ for $`\eta _{tot}=50`$. Fig. 6 The cross section of the system at $`t=3`$ for $`\eta _{tot}=50`$. Fig. 7 The cross section of the system at $`t=6`$ for $`\eta _{tot}=50`$. Fig. 8 Comparison of different models in the quasihomogeneous situation. Fig. 9 Relative error of the quasihomogeneous approach in the first model. Fig. 10 Relative error of the quasihomogeneous approach in the second model. Fig. 11 Relative error of the quasihomogeneous approach in the third model. Fig. 12 Relative error of the quasihomogeneous approach in the universal simulation.
warning/0003/hep-th0003125.html
ar5iv
text
# Untitled Document On non-$`L^2`$ solutions to the Seiberg–Witten equations by Christoph Adam, Bruno Muratori and Charles Nash Institut für Theoretische Physik Department of Mathematical Physics, Universität Karlsruhe, National University of Ireland, Germany. Maynooth, Ireland. Abstract: We show that a previous paper of Freund describing a solution to the Seiberg–Witten equations has a sign error rendering it a solution to a related but different set of equations. The non-$`L^2`$ nature of Freund’s solution is discussed and clarified and we also construct a whole class of solutions to the Seiberg–Witten equations. $`\mathrm{\S }`$ 1. Introduction With the introduction of the Seiberg–Witten equations there come a wealth of results on four manifold theory and a new improved point of view on Donaldson theory with an Abelian gauge theory supplanting a non-Abelian one—cf. for a review. An important vanishing theorem of , reminiscent of the Lichernowicz–Weitzenböck vanishing theorems, shows that there are no non-trivial solutions to the Seiberg–Witten equations on four manifolds with non-negative Riemannian scalar curvature. However one can have non-trivial solutions which are singular in some way—for example one could have a non-trivial solution which was not $`L^2`$: in Freund describes such a non-$`L^2`$ to the Seiberg–Witten equations on $`𝐑^4`$. Unfortunately a sign discrepancy in means that the expressions given there obey equations which differ from the Seiberg–Witten equations in a certain sign. These other equations also admit $`L^2`$ solutions as well as non-$`L^2`$ ones and so Freund’s equations are fundamentally different from the Seiberg–Witten equations. In $`\mathrm{\S }\mathrm{\hspace{0.17em}2}`$ we describe the Seiberg–Witten equations in a fairly explicit manner so as to expose notational conventions and matters of signs. In section $`\mathrm{\S }\mathrm{\hspace{0.17em}3}`$ we give the details concerning Freund’s work and then in section $`\mathrm{\S }\mathrm{\hspace{0.17em}4}`$ we give an $`L^2`$ solution of Freund’s equations and a class of solutions to the Seiberg–Witten equations. $`\mathrm{\S }`$ 2. The Seiberg–Witten equations If $`M`$ is an oriented Riemannian four manifold with metric $`g_{ij}`$ then the data we need for the Seiberg–Witten equations are a $`U(1)`$ connection $`A_i`$ on $`M`$ and a local spinor field $`M`$. If $`F_{ij}`$ is the curvature of $`A_i`$, so that its self-dual part $`F_{ij}^+`$ is given by $`F_{ij}^+=1/2(F_{ij}+(\sqrt{g}/2)ϵ_{ijkl}F^{kl})`$, then the Seiberg–Witten equations are $$\begin{array}{cc}\hfill F_{ij}^+& =\frac{i}{2}\overline{M}\mathrm{\Gamma }_{ij}M\hfill \\ \hfill \mathrm{\Gamma }^iD_iM& =0\hfill \end{array}$$ $`(2.1)`$ where $`\mathrm{\Gamma }_i`$ are the gamma matrices<sup>*</sup><sup>*</sup>Our conventions for the $`\mathrm{\Gamma }_i`$ are: $`\mathrm{\Gamma }_0=\left(\begin{array}{cc}0& I\\ I& 0\end{array}\right)`$, $`\mathrm{\Gamma }_i=\left(\begin{array}{cc}0& i\sigma ^i\\ i\sigma ^i& 0\end{array}\right)`$, and $`\mathrm{\Gamma }_5=\mathrm{\Gamma }_0\mathrm{\Gamma }_1\mathrm{\Gamma }_2\mathrm{\Gamma }_3`$ where $`\sigma ^i`$ are the Pauli matrices. satisfying $`\{\mathrm{\Gamma }_i,\mathrm{\Gamma }_j\}=2g_{ij}I`$, and $`D_i`$ and $`\mathrm{\Gamma }_{ij}`$ are given by $$D_i=_i+iA_i,\mathrm{\Gamma }_{ij}=\frac{1}{2}[\mathrm{\Gamma }_i,\mathrm{\Gamma }_j]$$ $`(2.2)`$ In Witten also quotes the equations using the two component spinor formalism of L. Witten and Penrose, cf. , where the Gamma matrices make no explicit appearance. In this spinor form the equations are $$\begin{array}{cc}\hfill _{A^{}B^{}}& =\frac{i}{2}\left(_A^{}\stackrel{~}{}_B^{}+_B^{}\stackrel{~}{}_A^{}\right)\hfill \\ \hfill 𝒟_{AA^{}}^A^{}& =0\hfill \end{array}$$ $`(2.3)`$ We now give a short summary of the relevant properties of the spinor formalism that we need here. With a Riemannian metric of signature $`(+,+,+,+)`$ the 4 components of a $`4`$-vector $`v_a(v_0,v_1,v_2,v_3)`$ are represented by a $`2\times 2`$ matrix which is denoted by $`v_{AA^{}}`$ and given by $$v_{AA^{}}=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}v_0+iv_3& iv_1+v_2\\ iv_1v_2& v_0iv_3\end{array}\right)$$ $`(2.4)`$ This expression for $`v_{AA^{}}`$ can be written as a linear combination of what are known as the Infeld–van der Waerden matrices $`g_{AA^{}}^a`$ defined by $$g_{AA^{}}^0=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right),g_{AA^{}}^i=\frac{i}{\sqrt{2}}\sigma ^i,i=1,2,3$$ $`(2.5)`$ where $`\sigma ^i`$ are the usual Pauli matrices so that $$\sigma ^1=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right),\sigma ^2=\left(\begin{array}{cc}0& i\\ i& 0\end{array}\right),\sigma ^3=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right)$$ $`(2.6)`$ Using the $`g_{AA^{}}^a`$’s the linear combination mentioned above is given by $$v_{AA^{}}=v_ag_{AA^{}}^a$$ $`(2.7)`$ and more generally if we have a tensor $`T_{a_1a_2\mathrm{}a_n}`$ it becomes $`𝒯_{A_1A_1^{}A_2A_2^{}\mathrm{}A_nA_n^{}}`$ where $$𝒯_{A_1A_1^{}A_2A_2^{}\mathrm{}A_nA_n^{}}=T_{a_1a_2\mathrm{}a_n}g_{A_1A_1^{}}^{a_1}g_{A_2A_2^{}}^{a_2}\mathrm{}g_{A_nA_n^{}}^{a_n}$$ $`(2.8)`$ In this formalism spinor indices are raised and lowered with the matrix $`ϵ_{AB}`$ defined by $$ϵ_{AB}=\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)=ϵ^{AB}$$ $`(2.9)`$ For example, for one spinor index, one can write $$v^A=ϵ^{AB}v_B,v_B=v^Aϵ_{AB}$$ $`(2.10)`$ An involution can also be defined on a spinor $`v^A`$ taking it to a spinor $`\stackrel{~}{v}^A`$ defined by $$v^A=\left(\begin{array}{c}\alpha \\ \beta \end{array}\right),\stackrel{~}{v}^A=\left(\begin{array}{c}\overline{\beta }\\ \overline{\alpha }\end{array}\right)$$ $`(2.11)`$ where bar means complex conjugate. Simplification occurs if the tensor is antisymmetric such as the curvature tensor $`F_{ij}`$: in that case one can verify that its spinor version $`_{AA^{}BB^{}}`$ is a linear combination of $`ϵ_{A^{}B^{}}`$ and $`ϵ_{AB}`$. More precisely one finds that $$F_{ij}g_{AA^{}}^ig_{BB^{}}^j=_{AA^{}BB^{}}=_{AB}ϵ_{A^{}B^{}}+_{A^{}B^{}}ϵ_{AB}$$ $`(2.12)`$ Moreover it also turns out that $`_{A^{}B^{}}`$ and $`_{AB}`$ are the spinor projections of the self-dual and anti-self-dual parts of $`F_{ij}`$ respectively; i.e. one can check that $$F_{ij}^+g_{AA^{}}^ig_{BB^{}}^j=_{A^{}B^{}}ϵ_{AB},F_{ij}^{}g_{AA^{}}^ig_{BB^{}}^j=_{AB}ϵ_{A^{}B^{}}$$ $`(2.13)`$ Now we return to the Seiberg–Witten equations and carry out the translation from the conventional to the spinor form. Starting with the Dirac equation we write $$M=\left(\begin{array}{c}\alpha \\ \beta \\ 0\\ 0\end{array}\right)\left(\begin{array}{c}^A^{}\\ 0\\ 0\end{array}\right)$$ $`(2.14)`$ and then the Dirac equation $$\mathrm{\Gamma }^iD_iM=0$$ $`(2.15)`$ becomes $$\sqrt{2}g_{AA^{}}^iD_i^A^{}=0$$ $`(2.16)`$ which we rewrite as $$𝒟_{AA^{}}^A^{}=0,\text{ with }𝒟_{AA^{}}=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}D_0+iD_3& iD_1+D_2\\ iD_1D_2& D_0iD_3\end{array}\right)$$ $`(2.17)`$ which is the desired form. Moving on to the other equation $$F_{ij}^+=\frac{i}{2}\overline{M}\mathrm{\Gamma }_{ij}M$$ $`(2.18)`$ we first display this equation in full as $$\begin{array}{cc}& \frac{1}{2}\left(\begin{array}{cccc}0& F_{01}+F_{23}& F_{02}F_{13}& F_{03}+F_{12}\\ F_{01}F_{23}& 0& F_{03}+F_{12}& F_{13}F_{02}\\ F_{13}F_{02}& F_{12}F_{03}& 0& F_{01}+F_{23}\\ F_{12}F_{03}& F_{02}F_{13}& F_{01}F_{23}& 0\end{array}\right)=\hfill \\ & \\ & \frac{1}{2}\left(\begin{array}{cccc}0& \overline{\beta }\alpha +\overline{\alpha }\beta & i\overline{\beta }\alpha i\overline{\alpha }\beta & |\alpha |^2|\beta |^2\\ \overline{\beta }\alpha \overline{\alpha }\beta & 0& |\alpha |^2|\beta |^2& i\overline{\beta }\alpha +i\overline{\alpha }\beta \\ i\overline{\beta }\alpha +i\overline{\alpha }\beta & |\alpha |^2+|\beta |^2& 0& \overline{\beta }\alpha +\overline{\alpha }\beta \\ |\alpha |^2+|\beta |^2& i\overline{\beta }\alpha i\overline{\alpha }\beta & \overline{\beta }\alpha \overline{\alpha }\beta & 0\end{array}\right)\hfill \end{array}$$ $`(2.19)`$ and then find that $$F_{ij}^+g_{AA^{}}^ig_{BB^{}}^j=\frac{i}{2}\overline{M}\mathrm{\Gamma }_{ij}Mg_{AA^{}}^ig_{BB^{}}^j$$ $`(2.20)`$ becomes $$_{A^{}B^{}}ϵ_{AB}=𝒯_{A^{}B^{}}ϵ_{AB}$$ $`(2.21)`$ i.e. $$_{A^{}B^{}}=𝒯_{A^{}B^{}}$$ $`(2.22)`$ where $$_{A^{}B^{}}=\frac{1}{2}\left(\begin{array}{cc}iF_{01}+iF_{23}+F_{13}F_{02}& iF_{12}iF_{03}\\ iF_{12}iF_{03}& iF_{01}iF_{23}+F_{13}F_{02}\end{array}\right)$$ $`(2.23)`$ and $$𝒯_{A^{}B^{}}=\frac{1}{2}\left(\begin{array}{cc}2i\overline{\alpha }\beta & i|\alpha |^2+i|\beta |^2\\ i|\alpha |^2+i|\beta |^2& 2i\alpha \overline{\beta }\end{array}\right)$$ $`(2.24)`$ On can now readily inspect equations 2.19 and 2.22, 2.23, 2.24 and confirm that the conventional and the spinor form of the equations agree. Also we can compute the matrix of components $`(i/2)(_A^{}\stackrel{~}{}_B^{}+_B^{}\stackrel{~}{}_A^{})`$ and verify that it is equal to $`𝒯_{A^{}B^{}}`$. Doing this we find that, if we start with $`^A^{}=(\alpha ,\beta )`$ and use 2.10 and 2.11, we obtain $$\begin{array}{cc}\hfill \frac{i}{2}((_A^{}\stackrel{~}{}_B^{}+_B^{}\stackrel{~}{}_A^{})& =\frac{1}{2}\left(\begin{array}{cc}2i\overline{\alpha }\beta & i|\alpha |^2+i|\beta |^2\\ i|\alpha |^2+i|\beta |^2& 2i\alpha \overline{\beta }\end{array}\right)\hfill \\ & =𝒯_{A^{}B^{}},\hfill \end{array}$$ $`(2.25)`$ as it should. We now turn to the explicit Fermion and gauge field considered by Freund. $`\mathrm{\S }`$ 3. Freund’s equations In Freund chooses $$\begin{array}{cc}\hfill A_i=\frac{1}{2r(rz)}\left(\begin{array}{c}0\\ y\\ x\\ 0\end{array}\right),\text{ and }^A^{}& =\frac{1}{2r\sqrt{r(rz)}}\left(\begin{array}{c}xiy\\ rz\end{array}\right)\hfill \\ \hfill \stackrel{~}{}^A^{}& =\frac{1}{2r\sqrt{r(rz)}}\left(\begin{array}{c}(rz)\\ x+iy\end{array}\right)\hfill \end{array}$$ $`(3.1)`$ for which one readily verifies that $$𝒟_{AA^{}}^A^{}=0$$ $`(3.2)`$ so that $`^A^{}`$ is indeed a zero mode. To check the other equation we compute the curvature and find that, if $`F_{ij}=_iA_j_jA_i`$, one has $$\begin{array}{cc}\hfill F_{0i}& =0,F_{12}=\frac{z}{2r^3},F_{13}=\frac{y}{2r^3},F_{23}=\frac{x}{2r^3}\hfill \\ \hfill _{A{}_{}{}^{}B_{}^{}}& =\frac{1}{4r^3}\left(\begin{array}{cc}yix& iz\\ iz& y+ix\end{array}\right)\hfill \end{array}$$ $`(3.3)`$ On the other hand one also finds that $$\begin{array}{cc}\hfill \frac{i}{2}(_A^{}\stackrel{~}{}_B^{}+_B^{}\stackrel{~}{}_A^{})& =\frac{1}{4r^3}\left(\begin{array}{cc}y+ix& iz\\ iz& yix\end{array}\right)\hfill \\ & =_{A{}_{}{}^{}B_{}^{}}\hfill \end{array}$$ so that $`_{A{}_{}{}^{}B_{}^{}}(i/2)(_A^{}\stackrel{~}{}_B^{}+_B^{}\stackrel{~}{}_A^{})`$ and Freund’s equations are $$\begin{array}{cc}\hfill _{A^{}B^{}}& =\frac{i}{2}\left(_A^{}\stackrel{~}{}_B^{}+_B^{}\stackrel{~}{}_A^{}\right)\hfill \\ \hfill 𝒟_{AA^{}}^A^{}& =0\hfill \end{array}$$ $`(3.4)`$ The Seiberg–Witten’s equations are known to admit no non-trivial regular $`L^2`$ solutions in flat space (or spaces of positive scalar curvature) so Freund was concerned to point out that his fields provide an example of a non-trivial solution which is not $`L^2`$. Unfortunately, as we have seen, Freund’s fields, though not $`L^2`$, are not solutions to the Seiberg–Witten equations. Since Freund’s fields are static and have a connection with $`A_0=0`$ it is natural to consider them in $`𝐑^3`$. We now do this letting $`𝐀=(A_1,A_2,A_3)`$ be the connection in $`𝐑^3`$ and denote its curvature components by $`\widehat{F}_{ab}`$, $`a,b=\mathrm{1..3}`$. We obtain thereby the three dimensional Freund equations $$\begin{array}{cc}\hfill \widehat{F}_{ab}& =ϵ_{abc}\overline{M}\sigma ^cM,a,b=\mathrm{1..3}\hfill \\ \hfill \text{/}_AM& =0,\text{where }\text{/}_A=i\sigma ^a(_a+iA_a)\hfill \end{array}$$ $`(3.5)`$ In similar fashion we could also have obtained the three dimensional Seiberg–Witten equations—cf. —and, as in four dimensions, these differ from Freund’s only in the sign of the quadratic Fermion term. They are $$\begin{array}{cc}\hfill \widehat{F}_{ab}& =ϵ_{abc}\overline{M}\sigma ^cM,a,b=\mathrm{1..3}\hfill \\ \hfill \text{/}_AM& =0,\text{where }\text{/}_A=i\sigma ^a(_a+iA_a)\hfill \end{array}$$ $`(3.6)`$ There is also a vanishing theorem which does not allow non trivial solutions in flat space so that there are no regular $`L^2`$ solutions to the equations 3.6 in $`𝐑^3`$. However there is no such restriction on the Freund’s equations 3.5. In the next section we show how to construct examples of singular non-$`L^2`$ solutions to the Seiberg–Witten equations and regular solutions to of Freund’s equations which are $`L^2`$ in $`𝐑^3`$. $`\mathrm{\S }`$ 4. The Freund and Seiberg–Witten equations in three dimensions First of all we simply note that Freund’s equations 3.5 (or indeed 3.4) admit the following regular solution which is $`L^2`$ in $`𝐑^3`$. $$\begin{array}{cc}\hfill ^A^{}& =\frac{\sqrt{12}(\mathrm{𝟏}+i\stackrel{}{\sigma }\stackrel{}{r})}{(1+r^2)^{3/2}}\left(\begin{array}{c}1\\ 0\end{array}\right)\hfill \\ \hfill A_i& =\frac{3}{(1+r^2)^2}\left(\begin{array}{c}2xz2y\\ 2yz+2x\\ 1r^2+2z^2\end{array}\right)\hfill \end{array}$$ $`(4.1)`$ as may be checked easily. Finally we would like to display some (necessarily singular) solutions to the three dimensional Seiberg–Witten equations 3.6; they will also of course be solutions of the full Seiberg–Witten equations 2.1 or 2.3. In fact we construct a whole class of such solutions parametrised by an arbitrary holomorphic function. First we need some facts about the Dirac equation in 3.6. A spinor $`M`$ that obeys the Dirac equation $`\text{/}_AM=0`$ of 3.6 must obey the condition $$_a\mathrm{\Sigma }^a=0,\text{where }\mathrm{\Sigma }^a=\overline{M}\sigma ^aM$$ $`(4.2)`$ The connection $`A_i`$ in the Dirac equation can be expressed in terms of the zero mode $`M`$ by writing $$\begin{array}{cc}\hfill A_i& =\frac{1}{\sqrt{\mathrm{\Sigma }^a\mathrm{\Sigma }^a}}\left(\frac{1}{2}ϵ_{ijk}_j\mathrm{\Sigma }_k+Im\overline{M}_iM\right)\hfill \\ & =\frac{1}{2}ϵ_{ijk}\left(_j\mathrm{ln}\sqrt{\mathrm{\Sigma }^a\mathrm{\Sigma }^a}\right)N_k\frac{1}{2}ϵ_{ijk}_jN_kIm\overline{\widehat{M}}_i\widehat{M}\hfill \\ \hfill \text{where }N^a& =\frac{\mathrm{\Sigma }^a}{\sqrt{\mathrm{\Sigma }^b\mathrm{\Sigma }^b}},\text{and }\widehat{M}=\frac{M}{\sqrt{\overline{M}M}}\hfill \end{array}$$ $`(4.3)`$ Now, if $`\chi `$ is the complex variable $$\chi =\frac{x+iy}{r^2}$$ $`(4.4)`$ and $`GG(\chi ,\overline{\chi })`$ is a function of $`\chi `$ and $`\overline{\chi }`$, we obtain a new class of zero modes $`M^G`$ where $$\begin{array}{cc}\hfill M^G& =e^{G/2}M^0\hfill \\ \hfill \text{and }M^0& =\frac{1}{r^3}\left(\begin{array}{c}z\\ x+iy\end{array}\right)\hfill \end{array}$$ $`(4.5)`$ The corresponding connection $`A_i^G`$ is found using formula 4.3 and is given by $$A_i^G=\frac{1}{2}ϵ_{ijk}_jGN_k$$ $`(4.6)`$ We note that the spinor $`M^0`$ is singular and non-$`L^2`$. For doing calculations it is also useful to note that $`M^0`$ is a solution of the free Dirac equation $$\text{/}M^0=0$$ $`(4.7)`$ and that the spin density for $`M^G`$ satisfies $$\begin{array}{cc}\hfill \overline{M^G}\sigma ^aM^G& =e^G\overline{M^0}\sigma ^aM^0\hfill \\ \hfill \text{where }\overline{M^0}\sigma ^aM^0& =\frac{N^a}{r^4}=\frac{1}{2i}(\overline{\chi })\times (\chi )\hfill \end{array}$$ $`(4.8)`$ The corresponding curvature $`\widehat{F}_{ij}^G`$ is $$\begin{array}{cc}\hfill \widehat{F}_{ij}^G& =\frac{ϵ_{ijk}}{2}[G_{,\chi }(\chi _{,kl}N_l+\chi _{,k}N_{l,l}\chi _{,ll}N_k\chi _{,l}N_{k,l})+\hfill \\ & G_{,\overline{\chi }}(\overline{\chi }_{,kl}N_l+\overline{\chi }_{,k}N_{l,l}\overline{\chi }_{,ll}N_k\overline{\chi }_{,l}N_{k,l})\hfill \\ & (G_{,\chi \chi }\chi _{,l}\chi _{,l}+G_{,\overline{\chi }\overline{\chi }}\overline{\chi }_{,l}\overline{\chi }_{,l}+2G_{,\chi \overline{\chi }}\chi _{,l}\overline{\chi }_{,l})N_k]\hfill \end{array}$$ $`(4.9)`$ After some tedious algebra we find that only the coefficient of $`G_{,\chi \overline{\chi }}`$ is non-zero and that $$\chi _{,k}\overline{\chi }_{,k}=\frac{2}{r^4}$$ $`(4.10)`$ and hence $$\widehat{F}_{ij}^G=\frac{2ϵ_{ijk}}{r^4}G_{,\chi \overline{\chi }}N_k$$ $`(4.11)`$ But to have a solution of the Seiberg–Witten equations we must require that $$\widehat{F}_{ij}^G=ϵ_{ijk}\overline{M^G}\sigma ^kM^G$$ $`(4.12)`$ and this means that $$\begin{array}{cc}\hfill \frac{2}{r^4}G_{,\chi \overline{\chi }}N_k& =\overline{M^G}\sigma ^kM^G\hfill \\ \hfill G_{,\chi \overline{\chi }}& =\frac{1}{2}e^G\hfill \end{array}$$ $`(4.13)`$ as may be easily checked. But this equation for $`G`$ is nothing other than the Liouville equation in the “target space coordinate” $`\chi `$ with the “wrong” sign; that is to say that the sign leads to the general singular solution $$G=\frac{|f^{}(\chi )|^2}{(1f\overline{f})^2}$$ $`(4.14)`$ where $`f`$ is an arbitrary holomorphic function of $`\chi `$. Hence any pair $`(M^G,A_i^G)`$ with $`G`$ given by 4.14 is a solution the Seiberg–Witten equations 3.6. In fact, these solutions resemble the two-dimensional solutions of the Seiberg–Witten equations that were discussed in . Their solutions emerged as solutions to the same Liouville equation 4.14, however the coordinate space variable $`x_+=x+iy`$ appears rather than the target space variable $`\chi `$ used here. Finally we want to briefly describe the geometry of the spin density term $`\overline{M}\sigma ^aM`$. It is clearly rotationally symmetric around the $`z`$ axis and the integral curves of $`\overline{M^0}\sigma ^aM^0`$ are circles that touch the $`z`$ axis at the point $`z=0`$. If we restrict to the $`xz`$ plane, then these integral curves are the field lines of a dipole in two dimensions, and the vector field $`\overline{M^0}\sigma ^aM^0`$ restricted to that plane is a scalar function times the field of a dipole in two dimensions. Acknowledgment: BM gratefully acknowledges financial support from the Training and Mobility of Researchers scheme (TMR no. ERBFMBICT983476). References 1.Witten E., Monopoles and four-manifolds, Math. Res. Lett., 1, 769–796, (1994). 2.Donaldson S. K., The Seiberg–Witten equations and 4-manifold topology, Bull. Amer. Math. Soc., 33, 45–70, (1996). 3.Freund P. G. O., Dirac monopoles and the Seiberg–Witten monopole equations, J. Math. Phys., 36, 2673–2674, (1995). 4.Penrose R. and Rindler W., Spinors and Space-time vol. 1, Cambridge University Press, (1984). 5.Loss M. and Yau H., Stability of Coulomb systems with magnetic fields III. Zero energy bound states of the Pauli operator, Commun. Math. Phys., 104, 283–290, (1986). 6.Nergiz S. and Sacioglu J., Liouville vortex and $`\varphi ^4`$ kink solutions of the Seiberg–Witten equations, J. Math. Phys., 37, 3753–3759, (1996).
warning/0003/hep-ph0003316.html
ar5iv
text
# INDIRECT CONSTRAINTS ON R-PARITY VIOLATING STOP COUPLINGS ## Acknowledgements The author thanks the organizers of the XXXV Rencontres de Moriond for the pleasant atmosphere in which this work was presented. Many thanks also go to Fabio Zwirner for helpful discussions and to Guido Martinelli for useful communications. ## Appendix We have calculated the contributions to $`K^0`$$`\overline{K}^0`$ mixing coming from the diagrams with $`R_p`$-violating (s)top couplings $`\lambda _{3jk}^{\prime \prime }`$. The calculation has been performed in the basis where the quark masses are diagonal, and all the flavor changing squark mass insertions have been neglected. We have also checked that in the minimal supergravity scenario considered in ref. $`^\mathrm{?}`$ the contributions coming from MSSM diagrams with quarks and charged higgs or squarks and charginos are negligible. The coefficients $`C_i`$ that appear in eq. (2) are: $`C_1`$ $`=`$ $`{\displaystyle \underset{i,j=1}{\overset{3}{}}}{\displaystyle \frac{g^4}{128\pi ^2}}K_{i1}^{}K_{i2}K_{j1}^{}K_{j2}m_{u_i}^2m_{u_j}^2[I_0+2I_2/m_W^2+I_4/4m_W^4](m_{u_i}^2,m_{u_j}^2,m_W^2,m_W^2)`$ $`\stackrel{~}{C}_1`$ $`=`$ $`{\displaystyle \underset{i,j=1}{\overset{2}{}}}{\displaystyle \frac{1}{4\pi ^2}}|\lambda _{313}^{\prime \prime }\lambda _{323}^{\prime \prime }|^2[(O_{i2}^tO_{j2}^t)^2I_4(m_b^2,m_b^2,m_{\stackrel{~}{t}_i}^2,m_{\stackrel{~}{t}_j}^2)+(O_{i2}^bO_{j2}^b)^2I_4(m_{\stackrel{~}{b}_i}^2,m_{\stackrel{~}{b}_j}^2,m_t^2,m_t^2)]`$ $`C_5`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle \frac{g^2}{4\pi ^2}}\lambda _{313}^{\prime \prime }\lambda _{323}^{\prime \prime }(O_{i2}^b)^2K_{31}^{}K_{32}m_t^2[I_2(m_{\stackrel{~}{b}_i}^2,m_W^2,m_t^2,m_t^2)+`$ (5) $`{\displaystyle \frac{1}{4m_W^2}}I_4(m_{\stackrel{~}{b}_i}^2,m_W^2,m_t^2,m_t^2)+{\displaystyle \frac{1}{4m_W^2\mathrm{tan}^2\beta }}I_4(m_{\stackrel{~}{b}_i}^2,m_{H^+}^2,m_t^2,m_t^2)]`$ $`+`$ $`{\displaystyle \underset{i,j,k=1}{\overset{2}{}}}{\displaystyle \frac{g^2}{8\pi ^2}}\lambda _{313}^{\prime \prime }\lambda _{323}^{\prime \prime }O_{j2}^tO_{k2}^tK_{31}^{}K_{32}[V_{i1}^{}O_{j1}^t{\displaystyle \frac{m_t}{\sqrt{2}m_W\mathrm{sin}\beta }}V_{i2}^{}O_{j2}^t]\times `$ $`\left[V_{i1}O_{k1}^t{\displaystyle \frac{m_t}{\sqrt{2}m_W\mathrm{sin}\beta }}V_{i2}O_{k2}^t\right]I_4(m_b^2,m_{\stackrel{~}{\chi }_i^+}^2,m_{\stackrel{~}{t}_j}^2,m_{\stackrel{~}{t}_k}^2)`$ and $`C_4=C_5`$. In the above equations, $`K_{ij}`$ are the CKM matrix elements, $`O_{ij}^t`$ and $`O_{ij}^b`$ are the left-right mixing matrices of the stop and sbottom sectors, and $`V_{ij}`$ is the mixing matrix of positive charginos as defined in ref. $`^\mathrm{?}`$. The masses of the supersymmetric particles and the mixing angles at the electroweak scale have been calculated with ISAJET $`^\mathrm{?}`$, and the common scale $`M_S`$ has been chosen as the geometrical mean of squark and chargino masses. The functions $`I_n`$ result from integration over the Euclidean momentum $`\overline{k}`$ of the four particles circulating in the loop: $$I_n(m_1^2,m_2^2,m_3^2,m_4^2)=_0^{\mathrm{}}\frac{\overline{k}^nd\overline{k}^2}{(\overline{k}^2+m_1^2)(\overline{k}^2+m_2^2)(\overline{k}^2+m_3^2)(\overline{k}^2+m_4^2)}$$ (6) ## References
warning/0003/gr-qc0003098.html
ar5iv
text
# 1 Flatness Problem ## 1 Flatness Problem Albrecht and Magueijo proposed that a time varying speed of light c should not introduce changes in the curvature terms in the Einstein’s equations in the cosmological frame and that Einstein’s equations must still hold. Assuming that matter behaves as a perfect fluid, the equations of state can be written as, $`P=(\gamma 1)\rho c^2(t).`$ (1) Friedmann equations for a homogeneous space time, with c and G, as functions of time are, $`{\displaystyle \frac{\dot{a}^2}{a^2}}={\displaystyle \frac{8\pi }{3}}G(t)\rho {\displaystyle \frac{Kc^2(t)}{a^2}}`$ (2) $`\ddot{a}={\displaystyle \frac{4}{3}}\pi G(t)(\rho +{\displaystyle \frac{3p}{c^2(t)}})a`$ (3) where $`\rho `$ and p are density and pressure of the matter, and K is the metric curvature parameter. Combining Eq.(2) and Eq.(3), the generalized conservation equation is, $`\dot{\rho }+3{\displaystyle \frac{\dot{a}}{a}}(\rho +{\displaystyle \frac{p}{c^2}})=\rho {\displaystyle \frac{\dot{G}}{G}}+{\displaystyle \frac{3Kc\dot{c}}{4\pi Ga^2}}`$ (4) We assume the conservation of ordinary matter; ie., the left hand side of Eq.(4) is zero. Thus the variations in c and G are such that the right hand side of Eq.(4) is identically zero. Assuming $`\rho a^{3\gamma }`$ and $`c(t)=c_0a^n`$,where $`c_0`$ and n are constants. The solution of the right hand side of Eq.(4), for $`3\gamma +2n20`$ is $$G=\frac{3K(c_0)^2n}{4\pi \rho _0}\frac{a^{3\gamma +2n2}}{3\gamma +2n2}+B$$ (5) and for, $`3\gamma +2n=2`$ is $$G=\frac{3K(c_0)^2n}{4\pi \rho _0}\mathrm{ln}a+B$$ (6) where B is a constant of integration. Thus the Friedmann equation for the $`3\gamma +2n2`$ case becomes $`{\displaystyle \frac{\dot{a}^2}{a^2}}=B^{}a^{3\gamma }+{\displaystyle \frac{K(c_0)^2a^{2n2}(23\gamma )}{(3\gamma +2n2)}}`$ (7) where B’is a constant. The curvature term in Eq.(2) vanishes as the scale factor evolves, if $`\ddot{a}>0`$ ie., $`\rho +\frac{3p}{c^2}<0`$ . Also the Eq.(4) gives the solution, $`\rho a^{3\gamma }`$ for $`\ddot{G}=\ddot{c}=0`$ if $`\rho +\frac{p}{c^2}0`$ Using the equation of state, Eq.(1), these conditions imply $`0\gamma <\frac{2}{3}`$ The scale factor evolves as $`a(t)t^{\frac{2}{3}\gamma }`$ if $`\gamma >0`$ and $`a(t)\mathrm{exp}(H_0(t)`$ if $`\gamma =0`$ For $`\gamma <\frac{2}{3}`$, the requirement $`\rho +\frac{3p}{c^2}<0`$ implies $`p<\frac{1}{3}\rho c^2`$, ie., the curvature term will vanish at large a only if the matter stress is gravitationally repulsive. From Eq.(7) it can be seen that the flatness problem can be solved for, $`n{\displaystyle \frac{1}{2}}(23\gamma ).`$ (8) This is exactly the same inequality, which Barrow derived without assuming energy conservation of ordinary matter$`^{\text{[4]}}`$. ## 2 Lambda Problem To incorporate the cosmological constant term into the Friedmann equation, a vacuum stress is considered obeying the equation of state, $`p_\mathrm{\Lambda }=\rho _\mathrm{\Lambda }c^2,`$ (9) with $`\rho _\mathrm{\Lambda }={\displaystyle \frac{\mathrm{\Lambda }c^2}{8\pi G}}0.`$ (10) The Friedmann equation containing $`\rho _\mathrm{\Lambda }`$is, $`{\displaystyle \frac{\dot{a}^2}{a^2}}={\displaystyle \frac{8}{3}}\pi G(\rho +\rho _\mathrm{\Lambda }){\displaystyle \frac{Kc^2}{a^2}}.`$ (11) As the universe expands the term containing $`\rho _\mathrm{\Lambda }`$ should dominate. But observationally it is very small. This is the $`\mathrm{\Lambda }`$ problem. The conservation Eq.(4) generalised to include $`\rho _\mathrm{\Lambda }`$ is, $`\dot{\rho }+3{\displaystyle \frac{\dot{a}}{a}}[\rho +{\displaystyle \frac{p}{c^2}}]=\dot{\rho _\mathrm{\Lambda }}(\rho +\rho _\mathrm{\Lambda }){\displaystyle \frac{\dot{G}}{G}}+{\displaystyle \frac{3Kc\dot{c}}{4\pi Ga^2}}.`$ (12) We assume a form for the variation of G and $`\mathrm{\Lambda }`$ in terms of the scale factor as, $`G=G_0a^q`$ and $`\mathrm{\Lambda }=\mathrm{\Lambda }_0a^s`$ where $`\mathrm{\Lambda }_0`$,q, $`G_0`$ and s are constants. Now Eq.(12) can be written as, $`\dot{\rho }+3{\displaystyle \frac{\dot{a}}{a}}{\displaystyle \frac{p}{c^2}}+{\displaystyle \frac{\dot{\mathrm{\Lambda }}c^2}{8\pi G}}=(\rho +\rho _\mathrm{\Lambda }){\displaystyle \frac{\dot{G}}{G}}+{\displaystyle \frac{3Kc\dot{c}}{4\pi Gc^2}}{\displaystyle \frac{2\mathrm{\Lambda }c\dot{c}}{8\pi G}}+{\displaystyle \frac{\mathrm{\Lambda }c^2\dot{G}}{8\pi G^2}}`$ (13) Assuming conservation of ordinary matter, we put the left hand side and right hand side of Eq.(13) separately to zero. The underlying assumption being that the variations of $`\mathrm{\Lambda }`$, G and c are such that conservation of energy, as true with the non varying parameters, is still valid. The solution for left hand side of Eq.(13), being zero, is $`\rho _\mathrm{\Lambda }={\displaystyle \frac{\mathrm{\Lambda }_0s(c_0)^2}{8\pi G_0}}{\displaystyle \frac{a(2n+sq)}{(2n+3\gamma +sq)}}+Ba^{3\gamma }`$ (14) where B is a constant of integration. The right hand side of Eq.(13) can be written as, $`{\displaystyle \frac{\mathrm{\Lambda }_0(c_0)^2}{8\pi G_0}}[{\displaystyle \frac{sq}{2n+3\gamma +sq}}2n]a^{2n+sq1}+{\displaystyle \frac{3K(c_0)^2n}{4\pi G_0}}a^{2nq3}qBa^{3\gamma 1}=0`$ (15) A solution for this equation is $`{\displaystyle \frac{sq}{2n+3\gamma +sq}}=2n,`$ (16) $`2nq3=3\gamma 1,`$ (17) and $`{\displaystyle \frac{3K(c_0)^2n}{4\pi G_0}}=qB.`$ (18) Using Eq.(17), Eq.(16) can be written as, $`{\displaystyle \frac{qs}{s+2}}=2n.`$ (19) Then $`q=\frac{3K(c_0)^2n}{4\pi G_0B}q_0n`$ and $`s=\frac{4}{q_02}`$ . For the dust era ($`\gamma =1`$) Eq.(17) becomes $`n={\displaystyle \frac{1}{q_02}}or4n=s`$ (20) For n to be negative, as in the solution of the flatness problem, $`q_0<2`$. For the radiation dominated era ($`\gamma =\frac{4}{3}`$), Eq.(17) is $`n={\displaystyle \frac{2}{q_02}}or2n=s`$ (21) The Friedmann equation for a time varying cosmological parameter, Eq.(13), becomes, $`{\displaystyle \frac{\dot{a}^2}{a^2}}={\displaystyle \frac{\mathrm{\Lambda }_0(c_0)^2}{3}}{\displaystyle \frac{(2n+3\gamma q)}{(2n+3\gamma +sq)}}a^{2n+s}+{\displaystyle \frac{8}{3}}\pi BG_0a^{q3\gamma }K(c_0)^2a^{2n2}`$ (22) For the dust dominated universe, $`\gamma =1`$ , using Eq.(20) $`{\displaystyle \frac{\dot{a}^2}{a^2}}={\displaystyle \frac{2\mathrm{\Lambda }_0(c_0)^2}{3(4n+2)}}a^{6n}+{\displaystyle \frac{8\pi BG_0}{3}}a^{2n2}K(c_0)^2a^{2n2}`$ (23) For $`n<\frac{1}{2}`$ the cosmological term will go to zero at large times faster than the other two terms of Eq.(23). For $`n<1`$ the curvature term will tend to zero at large times. Thus for $`n<\frac{1}{2}`$ all the terms will vanish at large times solving both the cosmological and the flatness problems. For the radiation dominated universe ($`\gamma =\frac{4}{3}`$ ), using equation (21) $`{\displaystyle \frac{\dot{a}^2}{a^2}}={\displaystyle \frac{2\mathrm{\Lambda }_0(c_0)^2}{3(2n+2)}}a^{4n}+{\displaystyle \frac{8\pi BG_0}{3}}a^{2n2}K(c_0)^2a^{2n2}`$ (24) For $`n<1`$ the $`\mathrm{\Lambda }`$ term will go to zero faster than the other two terms of the Eq.(24). As in the dust universe $`n<1`$ makes the curvature term vanish at large times. Thus in the radiation dominated universe $`n<1`$ solves both the cosmological and flatness problems. We have solved the flatness and $`\mathrm{\Lambda }`$ problems in a Friedmann universe by assuming that the variation of c, G and $`\mathrm{\Lambda }`$ are such that the conservation of matter with fixed c, G and $`\mathrm{\Lambda }`$ still holds. In the absence of $`\mathrm{\Lambda }`$ the solution is the same as that without assuming energy conservation. With $`\mathrm{\Lambda }`$ we can still solve the cosmological problem but with different exponents. Thus it might be worth exploring a more fundamental theory that will allow the variation of the parameters c, G, $`\mathrm{\Lambda }`$ without violating energy conservation. ## Acknowledgements PG thanks CSIR, NewDelhi for a research fellowship, and GVV thanks Prof. N. Dadhich for useful discussions and IUCAA, Pune for allowing to use their facilities.
warning/0003/hep-th0003233.html
ar5iv
text
# 1 Introduction ## 1 Introduction Dualities in string theory connect seemingly different theories. One of the most unexpected predictions in this web of dualities is the appearance of 11–dimensional theory (M–theory) that manifests itself at low energies as 11–dimensional supergravity. The theory compactified on different spaces is dual to different string theories. Compactification on $`S^1/ZZ_2`$ gives the M–theoretic extension of the heterotic $`E_8\times E_8`$ string, i.e. a geometric picture of 11–dimensional spacetime with two 10–dimensional walls at the ends of a finite interval along eleventh dimension . While the supergravity multiplet (graviton, gravitino and antisymmetric tensor fields) can penetrate in the $`d=11`$ bulk, the two $`E_8`$ gauge supermultiplets are confined to the two walls, respectively. The requirement of anomaly cancellation gives the relation between the 11–dimensional Newton’s constant $`\kappa `$ and the 10–dimensional gauge coupling constant $`\lambda `$ – it was first obtained in with a numerical factor of 2 missing. The correct relation was first obtained in and subsequently by many authors both in the downstairs and in the upstairs approaches . However, as was noted in , the previous calculations in the upstairs approach were inconsistent because of incorrect taking into account the $`ZZ_2`$ symmetry. Rather surprisingly, the result of the upstairs calculation of gave different result than the downstairs calculation of . Another interesting example of an orbifold in M–theory compactification is $`T^5/ZZ_2`$ which is dual to string theory IIB compactified on $`K3`$. There are 32 fixed six–planes and it turns out that there must be 16 additional twisted sector multiplets living on these six–planes to cancel anomalies. In the presence of five–branes some of these tensor multiplets are transferred to the branes. In section 2 we analyze the anomaly cancellation in the upstairs approach for the compactification on $`S^1/ZZ_2`$ and the results show complete equivalence of the downstairs and the upstairs approaches. The gauge anomaly is restricted to the walls and there are no mixing terms between the walls so that the cancellation proceeds exactly as in the downstairs approach. The combination $`\lambda ^6/\kappa ^4`$ turns out to be exactly the same as in (while in ref. it differs by a factor of 3). In section 3 we include five–branes and derive the field configuration for which the total anomaly vanishes. In section 4 we explicitly construct a theory compactified on $`T^5/ZZ_2`$ and show that there exists in the upstairs approach a configuration of fields that is locally (at each fixed six–plane) anomaly–free. Section 5 presents some conclusions. ## 2 M–theory on $`S^1/ZZ_2`$ The low energy limit of M–theory compactified on $`S^1/ZZ_2`$ was given in . We will use in this paper slightly different normalization of the fields (the same as in ). The actions for the theory in the upstairs and downstairs approaches look the same but the range of integration and the normalizations are different. In the upstairs approach we have full circle (of length $`2\pi \rho `$) in the eleventh dimension with two branes located at $`x^{11}=0`$ and $`x^{11}=\pi \rho `$ and we impose $`ZZ_2`$ symmetry on fields (for example $`C_{ABC}`$ and $`G_{ABCD}`$ with $`A,B,C,D=1\mathrm{}10`$ are odd under $`ZZ_2`$ and therefore vanish on the walls while $`C_{11AB}`$ and $`G_{11ABC}`$ are even). The action reads (for simplicity we keep only bosonic terms) $$S=\frac{1}{2\overline{\kappa }^2}_{\overline{M}_{11}}\sqrt{g}\left[R+\frac{1}{48}G^2\right]\frac{1}{12\overline{\kappa }^2}_{\overline{M}_{11}}CGG\frac{1}{4\lambda ^2}\underset{i}{}_{M_{10}^i}\sqrt{g}F_i^2.$$ (1) In the downstairs approach we take an interval (of length $`\pi \rho `$) with the two walls on its ends and instead of the $`ZZ_2`$ symmetry we have to impose appropriate boundary conditions. The action is the same but we have to replace $`\overline{M}_{11}`$ by $`M_{11}`$ (full circle by the interval in the eleventh dimension) and the coupling constant $`\overline{\kappa }`$ by $`\kappa `$ where: $$\kappa ^2=\frac{1}{2}\overline{\kappa }^2,_{M_{11}}=\frac{1}{2}_{\overline{M}_{11}}.$$ (2) Anomaly cancellation requires modifications of the Bianchi identities that involve sources on the walls. We will solve below for general $`G`$ and $`C`$ that satisfy these identities and are well defined on the full circle (i.e. are periodic). Let us start with some definitions that will prove useful later on. We introduce a space $`\mathrm{\Phi }`$ of differentiable functions defined on the interval $`[0,\pi \rho ]`$ and satisfying $$g\mathrm{\Phi }:g(0)=1,g(\pi \rho )=0.$$ (3) For any $`g\mathrm{\Phi }`$ we define two periodic functions, $`f_g^{(1)}`$ and $`f_g^{(2)}`$, defined on the circle $`(\pi \rho ,\pi \rho ]`$: $`f_g^{(1)}(x^{11})`$ $`=`$ $`\mathrm{sgn}(x^{11})g(|x^{11}|),`$ $`f_g^{(2)}(x^{11})`$ $`=`$ $`\mathrm{sgn}(x^{11})(g(|x^{11}|)1),`$ (4) $`f_g^{(1)}(\pi \rho )`$ $`=`$ $`f_g^{(2)}(\pi \rho )=0.`$ These functions constitute the most general $`ZZ_2`$–odd primitives of delta functions located, respectively, at 0 and $`\pi \rho `$, having no other singularities and defined globally on the circle. The derivatives of these functions are: $$\frac{f_g^{(i)}}{x^{11}}=2\delta ^{(i)}+h_g$$ (5) where $`h_g(x^{11})=g^{}(|x^{11}|)`$ is regular everywhere and $`\delta ^{(i)}`$ is the Dirac delta function located at the $`i`$–th wall. Later we will use the same symbol to denote the corresponding one form. It is easy to prove that for any $`g`$ $$_{S_1}𝑑x^{11}h_g(x^{11})f_g^{(i)}(x^{11})f_g^{(j)}(x^{11})=\frac{1}{3}\delta _{ij}.$$ (6) Regularization of the delta function gives also: $`\delta ^{(i)}f_g^{(j)}f_g^{(k)}`$ $``$ $`{\displaystyle \frac{1}{3}}(\delta _{ij}\delta _{ik})\delta ^{(i)},`$ (7) $`\delta ^{(i)}f_g^{(j)}`$ $``$ $`0`$ (8) (in sense of distributions i.e. when integrated with regular functions). Having defined the primitives of delta functions we can now solve for $`G`$ in the modified Bianchi identities. They read $$dG=\gamma \underset{i}{}\delta ^{(i)}I_i$$ (9) where $`\gamma =(4\pi )^2\overline{\kappa }^2/\lambda ^2`$ and $$I_i=\frac{1}{4\pi ^2}\left(\mathrm{tr}F_i^2\frac{1}{2}\mathrm{tr}R^2\right)|_{M_{10}^i}.$$ (10) An integral of $`dG`$ over a $`C_4\times I`$ where the interval $`I`$ surrounds only one wall is not vanishing but the integral over the entire interval should vanish for $`G`$ to be well defined. Therefore the sum of $`I_1`$ and $`I_2`$ must be cohomologically trivial: $$_{C_4}(I_1+I_2)=0.$$ (11) Because of (11) there exists a form of $`G`$ that is explicitly well defined globally. We can write $`I_1`$ $`=`$ $`H_4+d\mathrm{\Omega }_1,`$ $`I_2`$ $`=`$ $`H_4+d\mathrm{\Omega }_2`$ (12) where $`H_4`$ is a harmonic 4-form and $`\mathrm{\Omega }_i`$ are 3–forms well defined globally. Then we can write the solution of (9) in the form $$G=d\stackrel{~}{C}\frac{\gamma }{2}\mathrm{sgn}(x^{11})H_4+\gamma \underset{i}{}\delta ^{(i)}\mathrm{\Omega }_i$$ (13) where each of the terms is well defined globally. Unfortunately, we do not know the explicit form of (12) and it is difficult to connect $`\stackrel{~}{C}`$ with the $`C`$ field from M–theory where $`G=dC`$ in the bulk. Therefore we will seek the solution of (9) in a different form which is only implicitly well defined globally. The solution to (9) satisfying $`G=dC`$ in the bulk is given by $$G=dC+\gamma \underset{i}{}\delta ^{(i)}\omega _i$$ (14) where $`I_i=d\omega _i`$ locally<sup>1</sup><sup>1</sup>1In general, $`\omega _i`$ is not well defined globally since $`I_i`$ can have contributions from harmonic forms ($`H_4`$ in (12)) but $`C`$ is likewise not well defined globally and the two contributions should cancel to produce well defined $`G`$ like in (13).. In order to analyze the consequences of $`ZZ_2`$ symmetry, we will take care to define all forms globally at least in the eleventh dimension. We expect $`G`$ to be a regular form and therefore we expect that singularities in $`\delta C`$ should cancel $`\delta ^{(i)}`$ in (14). Let us define a regular three–form $`C_{reg}`$ (depending on functions $`g_1`$ and $`g_2`$) by the relation $$C=C_{reg}\frac{\gamma }{2}\underset{i}{}f_{g_i}^{(i)}\omega _i.$$ (15) Substituting this form to (14) we get a regular (along $`x^{11}`$) form $$G=dC_{reg}\frac{\gamma }{2}\underset{i}{}h_{g_i}dx^{11}\omega _i\frac{\gamma }{2}\underset{i}{}f_{g_i}^{(i)}I_i$$ (16) where we used (5). Integration of Bianchi identities over $`I\times C_4`$ where $`I`$ is an interval along $`x^{11}`$ comprising only one wall shows that $`G`$ in (16) is globally well defined only if $$g_1(x^{11})=g_2(x^{11})=g(x^{11}).$$ (17) Hence, the same function has to be used in $`f_g^{(1)}`$ and $`f_g^{(2)}`$ (and consequently there is only one regular derivative $`h`$). Therefore, we will suppress subscript $`g`$ from now on. To calculate all contributions to the anomaly, we have first to find gauge variation of $`C`$. Starting from the condition $`\delta G=0`$ and using equations (14) and (15) we get $$\delta C_{reg}=dB_{reg}\frac{\gamma }{2}h\underset{i}{}dx^{11}\omega _i^1$$ (18) where $`\delta \omega _i=d\omega _i^1`$. Let us now calculate the anomaly in the upstairs approach. The contribution from the topological term in the action is equal to $$\delta S_{top}=\frac{1}{12\overline{\kappa }^2}\delta CGG.$$ (19) The relation $`G=dC`$ valid in the bulk requires $`C`$ (and not $`C_{reg}`$) in the above expression – we would break supersymmetry in the bulk otherwise. Let us note that the authors of used $`\delta C_{reg}`$ in (19) and not $`\delta C`$ and it is one of the reasons for the difference between the results of the present paper and that of <sup>2</sup><sup>2</sup>2In a linear function was chosen for $`g`$ but as we show in the present paper nothing really depends on the choice of $`g`$ as long as conditions (3) are satisfied.. Using the expressions (16) and (18) we have $`\delta C`$ $`=`$ $`dB_{reg}{\displaystyle \frac{\gamma }{2}}h{\displaystyle \underset{i}{}}dx^{11}\omega _i^1{\displaystyle \frac{\gamma }{2}}{\displaystyle \underset{i}{}}f^{(i)}d\omega _i^1,`$ (20) $`G`$ $`=`$ $`dC_{reg}{\displaystyle \frac{\gamma }{2}}h{\displaystyle \underset{i}{}}dx^{11}\omega _i{\displaystyle \frac{\gamma }{2}}{\displaystyle \underset{i}{}}f^{(i)}I_i.`$ (21) It is easy to show (see relation (8)) that the terms with $`dB_{reg}`$ and $`dC_{reg}`$ do not contribute to (19) due to regularity of $`B_{reg}`$ and $`C_{reg}`$ and the fact that $`G_{ABCD}|_i=0`$. Therefore we get $`\delta S_{top}`$ $`=`$ $`{\displaystyle \frac{1}{12\overline{\kappa }^2}}\left({\displaystyle \frac{\gamma }{2}}\right)^3{\displaystyle _{\overline{M}_{11}}}{\displaystyle \underset{ijk}{}}\left(hdx^{11}\omega _i^1+f^{(i)}d\omega _i^1\right)`$ $`\left(hdx^{11}\omega _j+f^{(j)}I_j\right)\left(hdx^{11}\omega _k+f^{(k)}I_k\right)`$ $`=`$ $`{\displaystyle \frac{\gamma ^3}{96\overline{\kappa }^2}}{\displaystyle \underset{ijk}{}}({\displaystyle _{S_1}}hf^{(j)}f^{(k)}{\displaystyle _{M_{10}}}\omega _i^1I_jI_k`$ $`2{\displaystyle _{S_1}}hf^{(i)}f^{(k)}{\displaystyle _{M_{10}}}d\omega _i^1\omega _jI_k)`$ $`=`$ $`{\displaystyle \frac{\gamma ^3}{96\overline{\kappa }^2}}{\displaystyle \underset{ijk}{}}{\displaystyle _{M_{10}}}\left[\left({\displaystyle \frac{1}{3}}\delta _{jk}\right)\omega _i^1I_jI_k+2\left({\displaystyle \frac{1}{3}}\delta _{ik}\right)\omega _i^1I_jI_k\right].`$ Expanding this result shows that the terms which mix contributions from different walls cancel and we have the final result $$\delta S_{top}=\frac{\gamma ^3}{48\overline{\kappa }^2}\underset{i}{}_{M_{10}}\omega _i^1I_iI_i.$$ (23) The next contribution to the anomaly is the Green–Schwarz term – we take it to be of the form<sup>3</sup><sup>3</sup>3Choosing instead $`GX_7`$ with $`X_8=dX_7`$ would give the same contribution to the anomaly expressed as 12–form but the anomaly cancellation in the 10–form would then require a local counterterm. $$S_{GS}=\frac{1}{\gamma }_{\overline{M}_{11}}CX_8.$$ (24) Calculating the gauge variation of this term using (20) we get $`\delta S_{GS}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle _{\overline{M}_{11}}}{\displaystyle \underset{i}{}}(hdx^{11}\omega _i^1+f^{(i)}d\omega _i^1)X_8`$ (25) $`=`$ $`{\displaystyle _{\overline{M}_{11}}}{\displaystyle \underset{i}{}}\delta ^{(i)}\omega _i^1X_8={\displaystyle \underset{i}{}}{\displaystyle _{M_{10}}}\omega _i^1X_{8,i}.`$ The third contribution to the anomaly is the one–loop result – it is given by $$\delta S_{1loop}=\frac{\pi }{3}\underset{i}{}_{M_{10}}\omega _i^1I_iI_i+\underset{i}{}_{M_{10}}\omega _i^1X_{8,i}.$$ (26) The requirement of vanishing of the total anomaly i.e the sum of (23), (25) and (26) gives $$\frac{\lambda ^6}{\overline{\kappa }^4}=\frac{(4\pi )^5}{4}.$$ (27) Translating this result to the downstairs language using (2) we get $$\frac{\lambda ^6}{\kappa ^4}=(4\pi )^5$$ (28) and this relation is exactly the same as in . ## 3 M–theory on $`S^1/ZZ_2`$ with five–branes Let us now include five–branes – possible non–perturbative objects in M–theory that couple magnetically to $`G`$ and act as sources for modified Bianchi identities. Their presence can in general modify the condition (11) and therefore require non–standard identification of gauge fields with the connection. We would like to show here that we can repeat the whole discussion in the presence of five–branes in the upstairs approach. Let us take into account the five–branes parallel to three space dimensions and two CY dimensions (it is possible that with the non-standard embedding it is no longer a Calabi–Yau space but let us leave this subtlety aside). These five–branes act then as additional sources in the Bianchi identity $$dG=\gamma \underset{i}{}\delta ^{(i)}I_i\gamma \underset{\alpha }{}\delta _5^{(\alpha )}$$ (29) where $`\delta _5^{(\alpha )}`$ are products of five delta one forms along four CY dimensions and $`x^{11}`$. The requirement that $`G`$ be globally defined gives $$_{C_4}\left(I_1+I_2\right)+[C_\alpha ]=0$$ (30) where $`[C_\alpha ]`$ is equal to the number of five–branes surrounded by a given cycle $`C_4`$. In analogy to the previous case let us introduce a $`ZZ_2`$ odd and periodic function with jumps at $`x^{11}=x_{\alpha ^+}>0`$ and $`x^{11}=x_{\alpha ^+}`$ (it necessarily has two jumps symetrically located around $`x^{11}=0`$ because of $`ZZ_2`$ antisymmetry): $$f^{(\alpha )}(x^{11})=\mathrm{sgn}(x^{11})[g(|x^{11}|)\theta (x_{\alpha ^+}^{11}|x^{11}|)]$$ (31) where $`\theta `$ is the Heaviside step function. Solving for $`G`$ gives in analogy to (16) $$G=dC_{reg}\frac{\gamma }{2}\underset{i}{}hdx^{11}\omega _i\frac{\gamma }{2}\underset{i}{}f^{(i)}I_i\gamma \underset{\alpha ^+}{}hdx^{11}\theta _3^{(\alpha )}\gamma \underset{\alpha ^+}{}f^{(\alpha )}dx^{11}\delta _4^{(\alpha )}$$ (32) where $`d\theta _3^{(\alpha )}=\delta _4^{(\alpha )}`$. Therefore, we have to sum in (32) only over positive $`x_\alpha `$ (denoted by $`\alpha ^+`$) since the branes at negative values of $`x^{11}`$ are automatically taken into account. A similar argument as before (integration over a cycle comprising only one brane) shows that the function $`g`$ in (31) is the same for all branes and also the same as the one defining $`f^{(1)}`$ and $`f^{(2)}`$. The condition (30) that $`G`$ be globally defined is also crucial for anomaly cancellation and for confining anomaly to the wall or brane. Let us now describe the anomaly cancellation in the case with five–branes. We will work with 8–forms (and not 6–forms) and will not keep track of possible local counterterms. The contribution to the anomaly from the one–loop result is $$\delta S^{1loop}=\underset{\alpha }{}X_{8,\alpha }.$$ (33) The contribution to the anomaly from the Green–Schwarz mechanism is $$\frac{1}{\gamma }\delta GX_7=\frac{1}{\gamma }GdX_6^1=\underset{\alpha }{}X_{6,\alpha }^1.$$ (34) The contribution (34) translated to the 8–form language cancels the contribution from (33). Therefore, we expect that the contribution from the topological term should vanish. Let us show that (under some condition) this indeed is the case. Since $`\delta _5`$ is invariant under gauge and gravitational transformations, $`\delta C`$ is the same as before (20): $`\delta C`$ $`=`$ $`dB_{reg}{\displaystyle \frac{\gamma }{2}}h{\displaystyle \underset{i}{}}dx^{11}\omega _i^1{\displaystyle \frac{\gamma }{2}}{\displaystyle \underset{i}{}}f^{(i)}d\omega _i^1`$ (35) $`=`$ $`d\left(B_{reg}{\displaystyle \frac{\gamma }{2}}{\displaystyle \underset{i}{}}f^{(i)}\omega _i^1\right)+\gamma {\displaystyle \underset{i}{}}\delta ^{(i)}\omega _i^1.`$ Using (29) we get the five–brane contribution to the anomaly $$\delta CGG|_{fb}=2\gamma \underset{\alpha }{}\left(B_{reg}\frac{\gamma }{2}\underset{i}{}f^{(i)}\omega _i^1\right)\left(dC_{reg}\frac{\gamma }{2}\underset{j}{}f^{(j)}I_j\right)\delta _4^{(\alpha )}.$$ (36) Therefore, if we impose conditions for all positions of branes $`(x_\alpha ^{11})`$: $$B_{reg}(x_\alpha ^{11})=\frac{\gamma }{2}\underset{i}{}f^{(i)}(x_\alpha ^{11})\omega _i^1$$ (37) then the contribution to the anomaly from the topological term vanishes and summing up all contributions the total anomaly does so as well. ## 4 M–theory on $`T^5/ZZ_2`$ Another interesting orbifold compactification of M–theory is that on $`T^5/ZZ_2`$ i.e. down to a theory in six dimensions with 16 supercharges . Similarly as in the ten dimensional case there is a potential gravitational anomaly. In the limit of small compactification radius the particle content of this theory (after imposing $`ZZ_2`$ on the fields) is a chiral supergravity multiplet (consisting of a graviton, 2 chiral gravitinos and 5 selfdual twoforms) and five tensor multiplets (each consisting of an antiselfdual twoform, 2 antichiral fermions and 5 scalars). This theory does not possess vectors so that the potential anomaly must be purely gravitational. It turns out that the anomaly vanishes when there are additional 16 tensor multiplets (so called twisted sector not directly obtained from the compactification). $`T^5/ZZ_2`$ has 32 fixed six–planes and the additional matter multiplets should live on these six–planes. The anomaly from the small radius theory (untwisted sector only) is equally distributed among all 32 points so that the anomaly at the $`P`$-th plane is given by: $$I_{8,P}^{(1loop)}=\frac{1}{32}\left(2I^{(3/2)}10I^{(1/2)}\right)+N_P\left(2I^{(1/2)}I^{(3form)}\right)$$ (38) where $`N_P`$ is the number of the twisted sector matter multiplets living on the $`P`$-th fixed plane. Using explicit formulae for the anomalies in six dimensions one gets $$I_{8,P}^{(1loop)}=(N_P\frac{1}{2})X_8.$$ (39) Sum over $`P`$ gives the previously mentioned condition $`N_P=16`$. Since $`N_P`$ are integers it is impossible to cancel the anomaly at any fixed six-plane and to do so one has to modify Bianchi identities to provide additional source of anomaly inflow via Green–Schwarz mechanism. If $$dG=\underset{P}{}g_P\delta _5^{(P)}$$ (40) then by the same arguments as before the full anomaly is given by $$I_{8,P}=(N_Pg_P\frac{1}{2})X_8.$$ (41) To cancel the anomaly the magnetic charges have to be half–integers satisfying $$g_P=N_P\frac{1}{2}.$$ (42) We will describe below an explicit construction of $`G`$ satisfying (40) in the upstairs approach. We could apply the techniques worked out in the case of $`S^1/ZZ_2`$ but it is not necessary. The $`T^5/ZZ_2`$ orbifold is simpler because the cohomology condition (30) is now replaced by an algebraic relation $`N_P=16`$. Let us denote $`2^5=32`$ fixed six–planes by $`P(\overline{n})`$ where $`\overline{n}=(n_7,n_8,n_9,n_{10},n_{11})`$ and $`n_k=0,1`$. The action of $`ZZ_2`$ on the torus $`_k(\pi \rho _k,\pi \rho _k]`$ that leaves these 32 points intact is given by $`x_ix_i`$. Let us also introduce $`5\times 2^4=80`$ intervals joining fixed points by $`I(\overline{n}^k)`$ where $`\overline{n}^k=(n_7,\mathrm{},n_{11})`$ with $`n_k`$ left out. The interval $`I(\overline{n}^k)`$ is parallel to the $`k`$-th axis and we assume that its orientation is the same as the orientation of that axis. The field $`G`$ satisfying (40) and antisymmetric under $`ZZ_2`$ is given by $$G=dC+\frac{1}{2}\underset{k}{}\underset{\overline{n}^k}{}c(\overline{n}^k)\mathrm{sgn}(x^k)\delta _4^{(\overline{n}^k)}$$ (43) where $$\delta _4^{(\overline{n}^k)}=\frac{1}{4!}ϵ_{kpqrs}\delta \left(x^pn_p\pi \rho _p\right)\mathrm{}\delta \left(x^sn_s\pi \rho _s\right)$$ (44) and $`c(\overline{n}^k)`$ are constants characterizing $`G`$ along the interval $`I(\overline{n}^k)`$. Calculating $`dG`$ from (43) and comparing with (40) and (42) we get $$g_{P(\overline{n})}=\underset{k}{}(1)^{n_k}c(\overline{n}^k)=N_P\frac{1}{2}.$$ (45) In the above equation $`\overline{n}^k`$ has the same components as $`\overline{n}`$ but with $`n_k`$ dropped. It is straightforward for any given set of $`N_P`$ (satisfying $`N_P=16`$ i.e. $`g_P=0`$) to get $`c(\overline{n}^k)`$ satisfying the above equation. Let us describe explicitly for example the “checkerboard” configuration where 16 tensor multiplets are distributed in the most uniform way: $$N_P=\frac{1}{2}\left[1+(1)^{n_7+n_8+n_9+n_{10}+n_{11}}\right].$$ (46) Shrinking any of the radii in the above configuration, any pair of fixed points along the shrinking direction contributes 1 tensor multiplet what corresponds to the string limit . For this choice for $`N_P`$ it is easy to derive one of possible sets $`c(\overline{n}^k)`$: $$c(\overline{n}^k)=\frac{1}{10}(1)^{n_7+\mathrm{}n_{11}}$$ (47) (where the index $`n_k`$ is left out on the r.h.s.) and check that (45) is satisfied. There are of course other $`c(\overline{n}^k)`$ – differring by $`dC`$ in (43) – that give the same local cancellation of anomalies. Other distributions of tensor multiplets among the fixed planes can be obtained from the “checkerboard” configuration by moving those mutliplets one by one from some fixed point to the neighbouring one along interval $`I(\overline{n}^k)`$. Such change requires appropriate modification of $`c(\overline{n}^k)`$: $$\mathrm{\Delta }c(\overline{n}^k)=\pm \frac{1}{2}$$ (48) where the sign depends on the “orientation” of the exchange. The theory in the presence of five–branes orthogonal to $`T^5`$ can be similarly analyzed. The location of the $`\alpha `$-th five–brane is given by a vector with components $`x_\alpha ^k`$ for $`k=7,\mathrm{},11`$. In the presence of such five–branes the modified Bianchi identities read $$dG=\underset{P}{}g_P\delta _5^{(P)}+\underset{\alpha }{}\delta _5^{(\alpha )}.$$ (49) Because of the $`ZZ_2`$ symmetry the five–branes must come in pairs with opposite coordinates: $`x_\beta ^k=x_\alpha ^k`$. In order to satisfy (49) the field $`G`$ must have jumps at the location of both five–branes. The requirement that $`G`$ is globally well defined can be fulfilled only when there are two additional jumps of opposite sign located at some fixed six–planes. Such negative jumps of $`G`$ at the fixed planes correspond to removing two twisted sector matter multiplets from those fixed planes and transfering them to the so called “wandering” five–branes. One of the solutions is to put both of those negative jumps for all pairs of five–branes at the origin $`P(\overline{n}=(0,0,0,0,0))`$. If we want to transfer tensor multiplets from some other fixed six–planes we have first to move those multiplets to the origin using the procedure described before (48). The solution transferring all multiplets from the origin to the branes is given by the sum of (43) and $$\mathrm{\Delta }G=\frac{1}{25!}\underset{\alpha }{}\underset{pqrst}{}ϵ_{pqrst}\mathrm{sgn}(x^p)\theta (|x_\alpha ^p||x^p|)\delta \left(x^qx_\alpha ^q\right)\mathrm{}\delta \left(x^tx_\alpha ^t\right).$$ (50) Now 16 twisted sector tensor multiplets are distributed between fixed planes and five–branes $$\underset{P}{}N_P+\underset{\alpha }{}1=16.$$ (51) The extremal situation is reached when all tensor multiplets sit on five–branes and there are no twisted sector multiplets on the fixed six–planes. ## 5 Conlusions We have analyzed the problem of anomaly cancellation in the upstairs approach. It turned out that for the compactification on $`S^1/ZZ_2`$ orbifold all final results are sums of contributions from two walls so the anomaly cancellation can be achieved without any corrections to the original action. $`C`$ and $`G`$ fields depend on an arbitrary function of $`x^{11}`$ but the anomaly cancellation works in the same way for any choice of this function (however, one should note that for other purposes some choice may be preferred over the others like for example in where it was shown that the Kaluza–Klein zero modes have some specific dependence on the eleventh dimension). The cancellation of anomalies in the presence of five–branes is possible but requires one additional condition on the variation of $`C`$ under the gauge and gravitational transformations. In the case of compactification on $`T^5/ZZ_2`$ (with and without five–branes) we have presented in the upstairs approach an explicit method to derive field configuration for which the total anomaly vanishes separately at each fixed six–plane. The same results can be obtained in the downstairs approach but with appropriate boundary conditions replacing the $`ZZ_2`$ symmetry. This, however, is less straightforward to implement. For example the upstairs approach automatically takes into account proper normalization of charges. In the case of the $`T^5/ZZ_2`$ compactification the number of additional tensor multiplets is always 16 in the upstairs approach (since two $`ZZ_2`$ symmetric “wandering branes” carry two tensor multiplets) while in the downstairs approach it can be smaller (because now there is only a single brane with one tensor multiplet). In conclusion, from the point of view of anomaly cancellation, the upstairs and downstairs approaches are equivalent. In the actual computation, however, the upstairs approach seems to be more convenient. Acknowledgements We would like to thank J. Conrad and H.P. Nilles for discussions. K.A.M. was partially supported by the Polish KBN grant 2P03B 03715 (1998-2000). M.O. was partially supported by the Polish grant KBN 2 P03B 052 16 (1999-2000).
warning/0003/cs0003057.html
ar5iv
text
# XNMR – A tool for knowledge bases exploration ## General Information The XNMR package is an attempt of integration between the well-founded semantics system XSB(??) and the stable models evaluator SMODELS(?). It works in UNIX platforms (eg. Linux, Solaris), and work is being done to port it to Windows NT. The package consists of three layers. The bottom layer is a low-level interface between XSB and SMODELS. This interface is written in C, and consists of 177 lines of code. The second layer is a high-level library of Prolog predicates used to communicate with SMODELS. It has 117 lines of Prolog code. The third layer is the top-loop evaluator. It is also written in Prolog, and consists of 421 lines of code. ## Description of the System XNMR is a package intended to integrate the possible-worlds stable semantics of SMODELS with the more skeptical well-founded semantics of XSB. It is aimed at allowing the exploration of knowledge bases. It is well-suited for debugging large knowledge bases, by exploiting the modularity of such designs. The package uses XSB as a pre-processing phase, where the well-founded semantics of a program, with respect to a given query, is computed. As a result of this process, a residual program is computed, where the interdependencies of the undefined objects is represented. The residual program is, then, given as input to SMODELS, so that the stable models for these undefined objects can be computed. ## Applying the System ### Methodology This work is exploratory, in the sense that it combines characteristics of two well known systems for logic programming in a general way. We believe the system may be applied to several situations where the stable models semantics is preferred, as long as the notion of relevancy to a query (see Specifics) is satisfied. More specifically, one application of this system is in debugging large knowledge bases. The system can take advantage of its ability to deal with partial information to allow for debugging of separate parts, or modules, of a knowledge base. Work is in progress to develop a programming methodology that will allow problems to be described in such a way as to be suitable for evaluation with XNMR. The methodology is based on ways to encode the problem such as to avoid the non-relevancy problem. ### Specifics XNMR computes the partial stable models which are total on the relevant program, given a query. This is due to the combination of a query-driven system (XSB) to a bottom-up evaluator like SMODELS. The models computed are total stable models with respect to the relevant parts of the program, according to the given query. We believe this system may have successful applications on areas like debugging of knowledge bases, and null-valued databases. ### Users and Useability The system is to be used by people with an understanding of the issues involved in the combination of well-founded and stable models semantics. Since it is integrated with XSB Prolog, it is general enough to be used in several applications. Even though the system is not, at the moment, being used outside our research group, there are already people interested in applying it to several applications. ## Evaluating the System ### Benchmarks To our knowledge, there are no comparative systems available today. One reasonable way to benchmark the systems, though, would be considering XSB as a pre-processor to SMODELS. This way, accepted benchmarks for SMODELS could be applied to the system, as long as they were well-behaved with respect to the relevant-program restriction imposed by the top-down nature of XSB evaluation. The system is incorporated to the XSB Prolog system, therefore having the same user-friendliness associated to such systems. ### Comparison The system is inherently logic-based. As stated in the previous section, there are no standard benchmarks to the system, due to its unique combination of top-down and bottom-up evaluators, which results in a restricted semantics being computed. We believe such benchmarks should be considered in an applications-based framework. ### Problem Size The XNMR system, by itself, doesn’t impose any restrictions to the size of problems being handled. So, the restrictions of the system are those of XSB and SMODELS. We also note that the system is not a prototype, and has been fully implemented as a package of XSB.
warning/0003/cond-mat0003268.html
ar5iv
text
# From second to first order transitions in a disordered quantum magnet \[ ## Abstract We study the spin-glass transition in a disordered quantum model. There is a region in the phase diagram where quantum effects are small and the phase transition is second order, as in the classical case. In another region, quantum fluctuations drive the transition first order. Across the first order line the susceptibility is discontinuous and shows hysteresis. Our findings reproduce qualitatively observations on LiHo<sub>x</sub>Y<sub>1-x</sub>F<sub>4</sub>. We also discuss a marginally stable spin-glass state and derive some results previously obtained from the real-time dynamics of the model coupled to a bath. preprint: LPTENS-97/18 \] The study of quantum effects on the properties of spin glasses is a subject of great experimental and theoretical interest. Spin-glass phases have been identified in systems such as mixed hydrogen-bonded ferroelectrics , the dipolar magnet LiHo<sub>x</sub>Y<sub>1-x</sub>F<sub>4</sub> or Sr-doped La<sub>2</sub>CuO<sub>4</sub> where quantum mechanics plays a fundamental role. An important question is whether quantum spin glasses are qualitatively different from their classical counterparts at low temperature. There is growing experimental evidence that the answer to this question is affirmative both in and out of equilibrium . The thermodynamics of several models of disordered magnetic systems has been investigated with various techniques. Mean-field-like models have been solved using the replica formalism in imaginary-time and the Ising model in a transverse field has also been studied in finite dimensions. It is generally found that, in terms of a suitably defined quantum parameter $`\mathrm{\Gamma }`$, a boundary $`\mathrm{\Gamma }_\mathrm{c}(T)`$ in the $`\mathrm{\Gamma }T`$ plane separates spin-glass (SG) and paramagnetic (PM) phases. The transition line ends at a quantum critical point at $`T=0,\mathrm{\Gamma }_\mathrm{c}(0)`$ above which the system is paramagnetic at all temperatures. In the case of the quantum spherical $`p`$-spin model, the real-time dynamics of the system coupled to a phonon bath was also investigated . In this case, a boundary $`\mathrm{\Gamma }_\mathrm{d}(T)`$ was found across which there is a dynamic phase transition from a PM state with equilibrium dynamics to a SG with non-stationary, aging, dynamics. In this paper we investigate in detail the equilibrium properties of this model. We find that a tricritical point $`(T^{},\mathrm{\Gamma }^{})`$ divides the line $`\mathrm{\Gamma }_\mathrm{c}(T)`$ in two parts. For $`TT^{}`$, the SG transition is of second order and the behavior of the quantum system is qualitatively similar to that of the classical one. However, for $`T<T^{}`$ quantum fluctuations drive the transition first order. The magnetic susceptibility is discontinuous and shows hysteresis across the first-order line. These findings reproduce qualitatively the observed behavior of LiHo<sub>x</sub>Y<sub>1-x</sub>F<sub>4</sub> in a transverse magnetic field . The equations describing this system are non-linear and there is multiplicity of solutions in parts of the phase diagram. We found as a surprise that the usual criteria used to choose between them have to be reinterpreted in order to get physically meaningful solutions in the region $`T<T^{}`$. We also discuss the properties of solutions obtained through the use of the marginality condition, an approach recently applied to the study of quantum problems. It is known from work on classical models that the results of this approach are closely related to those obtained from the analysis of the real-time dynamics of the system. We explicitly show that this holds true in our quantum case. This unables us to identify the dynamical transition line $`\mathrm{\Gamma }_\mathrm{d}(T)`$ and to derive certain properties of the non-equilibrium dynamics using the replica calculation. The Hamiltonian of the quantum $`p`$-spin spherical model is $$H[𝐏,𝐬,J]=\frac{1}{2M}\underset{i=1}{\overset{N}{}}P_i^2\underset{i_1<\mathrm{}<i_p}{\overset{N}{}}J_{i_1\mathrm{}i_p}s_{i_1}\mathrm{}s_{i_p},$$ (1) where $`s_i`$ is a scalar spin variable and the conjugated momenta $`P_i`$ satisfy the commutation relations $`[P_i,s_j]=i\mathrm{}\delta _{ij}`$. A Lagrange multiplier $`z`$ enforces the spherical constraint $`1/N_{i=1}^Ns_i^2=1`$. The interactions $`J_{i_1\mathrm{}i_p}`$ are chosen from a Gaussian distribution with zero mean and variance $`[J_{i_1\mathrm{}i_p}^2]_J=\stackrel{~}{J}^2p!/(2N^{p1})`$. The model has glassy properties for all $`p2`$. The Hamiltonian (1) may be interpreted in several ways. It represents a non-linear generalization of the quantum-rotor spin-glass models discussed in the literature . It also describes a quantum particle moving in an $`N`$ (eventually infinite) dimensional space in the presence of a random potential. Finally, its partition function is formally identical to that of a classical chain of “length” $`L=\beta \mathrm{}`$ embedded in an $`N`$-dimensional random environment. The equilibrium properties of the model are obtained using a replicated imaginary-time path integral formalism . In the large $`N`$ limit, the saddle-point evaluation of the partition-function allows us to define the order-parameter $`Q_{ab}(\tau \tau ^{})=1/N_{i=1}^N𝒯s_i^a(\tau )s_i^b(\tau ^{})`$ where $`a,b=1,\mathrm{},n`$ denote the replica indices and $`𝒯`$ the imaginary-time ordering operator. In terms of $`Q_{ab}`$ the free-energy per spin reads $`\mathrm{F}`$ $`=`$ $`\underset{n0}{lim}{\displaystyle \frac{1}{2n}}\{{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{\omega _k}{}}[\text{Tr}\mathrm{ln}\left({\displaystyle \frac{\stackrel{~}{Q}(\omega _k)}{\beta \mathrm{}}}\right)n((M\omega _k^2+z)`$ (2) $`\times `$ $`{\displaystyle \frac{\stackrel{~}{q}_d(\omega _k)}{\mathrm{}}}1)]nz{\displaystyle \frac{\stackrel{~}{J}^2}{2\mathrm{}}}{\displaystyle \underset{ab}{}}{\displaystyle _0^\beta \mathrm{}}d\tau Q_{ab}^p(\tau )\}`$ (3) where $`\beta =1/(k_BT)`$ is the inverse temperature, $`\omega _k=2\pi k/\beta `$ are the Matsubara frequencies, $`\stackrel{~}{Q}_{ab}(\omega _k)=_0^\beta \mathrm{}𝑑\tau Q_{ab}(\tau )e^{i\omega _k\tau }`$ and $`\stackrel{~}{q}_d(\omega _k)=\stackrel{~}{Q}_{aa}(\omega _k)`$. ¿From here on we take $`\stackrel{~}{J}`$ as the unit of energy, $`\mathrm{}/\stackrel{~}{J}`$ as the unit of time, and work with dimensionless quantities. Quantum fluctuations are controlled by the parameter $`\mathrm{\Gamma }\mathrm{}^2/(\stackrel{~}{J}M)`$. The classical limit of the model is recovered when $`\mathrm{\Gamma }0`$. The equilibrium solutions are determined by requiring that $`\stackrel{~}{Q}_{ab}(\omega _k)`$, parametrized according to different ansatze, be an extremum of $`\mathrm{F}`$. In the following we concentrate on the case $`p3`$. The phenomenology of the $`p=2`$ case is not as rich. For sufficiently high $`T`$ and/or $`\mathrm{\Gamma }`$, thermal and/or quantum fluctuations destroy the SG phase and the system is in the PM phase. The free-energy is then extremal for $`Q_{ab}(\tau )=\delta _{ab}q_d(\tau )`$. Its Fourier transform is the solution of the equation $`\stackrel{~}{q_d}(\omega _k)=\left[\omega _k^2/\mathrm{\Gamma }+z\mathrm{\Sigma }(\omega _k)\right]^1,`$ (4) with $`\mathrm{\Sigma }(\omega _k)=p/2_0^\beta 𝑑\tau q_d^{p1}(\tau )e^{i\omega _k\tau }`$ and $`z`$ is determined from $`q_d(0)=1`$. The above equation is non-linear and may have several solutions, some of which may be spurious. We solved Eq. (4) numerically for $`p=3`$. We found that for $`T>T^{}1/6`$ there is only one solution, irrespective of the value of $`\mathrm{\Gamma }`$. However, for $`T<T^{}`$, three solutions coexist in a finite region of the $`T\mathrm{\Gamma }`$ plane (not including the $`\mathrm{\Gamma }`$=0 axis). One of them is unstable and can be discarded from the start. We discuss below how to choose the physical solution between the remaining two. In the SG phase, inspired by the classical case , we searched for one-step RSB solutions of the form $`Q_{ab}(\tau )=q_d^{}(\tau )\delta _{ab}+q_{\mathrm{EA}}ϵ_{ab}`$, where $`ϵ_{ab}=1`$ if $`a`$ and $`b`$ belong to the same diagonal block of size $`m\times m`$ and zero otherwise, and $`q_d^{}(\tau )=q_d(\tau )q_{\mathrm{EA}}`$. The diagonal part, $`q_d(\tau )`$, the breaking point, $`m`$, and the Edwards-Anderson order parameter, $`q_{\mathrm{EA}}`$, are determined by extremizing $`\mathrm{F}`$. We find $$\stackrel{~}{q}_d^{}(\omega _k)=\left[\omega _k^2/\mathrm{\Gamma }+z^{}(\mathrm{\Sigma }^{}(\omega _k)\mathrm{\Sigma }^{}(0))\right]^1,$$ (5) where $`\mathrm{\Sigma }^{}(\tau )=p/2(q^{p1}(\tau )q_{\mathrm{EA}}^{p1})`$, $`z^{}=p/2\beta mq_{\mathrm{EA}}^{p1}(1+x_p)/x_p`$ and $$m=Tx_p\sqrt{2/(p(1+x_p))}q_{\mathrm{EA}}^{p/2}.$$ (6) The parameter $`x_p`$, solution of an algebraic equation, depends on $`p`$ only, and we found $`x_3=1.817`$. The condition $`q_d(0)=1`$ now yields an equation for the breaking point of the form $`m\mu _T(\mathrm{\Gamma })`$. Solutions of Eq. (5) exist only for $`\mathrm{\Gamma }\mathrm{\Gamma }_{\mathrm{max}}(T)`$. Above this value, quantum fluctuations destroy the SG phase. We found that the function $`\mu _T(\mathrm{\Gamma })`$ has two real branches in the interval $`0\mathrm{\Gamma }\mathrm{\Gamma }_{\mathrm{max}}(T)`$. The physical values of $`m`$ are on the lowest branch which verifies $`\mu _T(0)=m_{\mathrm{class}}(T)`$, the classical breaking point parameter. The situations above and below $`T^{}`$ are different. For $`TT^{}`$ (but lower than the classical transition temperature), $`m_{\mathrm{max}}m(\mathrm{\Gamma }_{\mathrm{max}})=1`$, its largest possible value. For $`T<T^{}`$, instead, $`m_{\mathrm{max}}<1`$. In both cases, $`q_{\mathrm{EA}}`$ is finite at $`\mathrm{\Gamma }_{\mathrm{max}}(T)`$. We discuss below the consequences of these facts. We found that $`lim_{T0}m_{\mathrm{max}}(T)=0`$, implying that replica symmetry is restored at the quantum critical point as in the model discussed in reference . All these conclusions, obtained from the numerical analysis of the $`p=3`$ case, also follow from an approximate analytical solution of the equations for arbitrary $`p3`$ . PM solutions exist throughout the $`T\mathrm{\Gamma }`$ plane. The free-energies of the different states must thus be compared in order to construct a phase diagram. Figs. 1(a) and (b) show the $`\mathrm{\Gamma }`$-dependence of the PM and SG free-energies for the case $`p=3`$ computed from Eq. (3) for two temperatures, above and below $`T^{}`$. Solid lines and symbols represent the PM and SG solutions, respectively. The curves end at the point where the corresponding solution disappears. It may be seen that for $`T>T^{}`$ the free-energies of the two states intersect precisely at $`\mathrm{\Gamma }_\mathrm{c}(T)=\mathrm{\Gamma }_{\mathrm{max}}(T)`$: the SG solution does not extend beyond the transition point and no hysteresis is expected. Below the critical point, $`\mathrm{F}_{\mathrm{SG}}>\mathrm{F}_{PM}`$ meaning that the SG solution maximizes the free-energy. This is the usual situation encountered in replica theories of classical spin glasses. As in the classical case, $`q_{\mathrm{EA}}`$ is discontinuous at the transition. The latter is nevertheless of second order because $`m=1`$ at $`\mathrm{\Gamma }_\mathrm{c}`$ and, therefore, the effective number of degrees of freedom participating in the transition $`(1m)q_{\mathrm{EA}}0`$ at $`\mathrm{\Gamma }_\mathrm{c}`$. There is no latent heat and the linear susceptibility is continuous. The situation is more involved below $`T^{}`$. To start with, one has to choose between the two PM solutions labeled PM<sub>1</sub> and PM<sub>2</sub> in Fig. 1. Naively, one would choose the solution with the lowest free-energy, i.e., PM<sub>2</sub>. However, this solution has unphysical properties. As shown in Fig. 1(b), its free-energy never intersects that of the SG phase. Both the free-energy and the susceptibility diverge as $`T0`$. Furthermore, this solution disappears at a finite value of $`\mathrm{\Gamma }`$ (not shown in the figure) and cannot thus be reached starting from $`\mathrm{\Gamma }=\mathrm{}`$. On the other hand, it is clear that the ground-state of Hamiltonian (1) must have finite susceptibility and energy. We thus conclude that PM<sub>2</sub> is a spurious solution and that PM<sub>1</sub> has to be chosen even if its free-energy is higher . The free-energies of the SG and PM<sub>1</sub> states cross at $`\mathrm{\Gamma }_\mathrm{c}<\mathrm{\Gamma }_{\mathrm{max}}`$ as shown in the inset in Fig. 1(b). In the low temperature phase, $`\mathrm{F}_{\mathrm{SG}}<\mathrm{F}_{\mathrm{PM}}`$, the opposite of what we found for $`T>T^{}`$. The SG and PM solutions extend beyond the point where they cross. There is a region of phase coexistence and hysteresis effects are thus expected in the behavior of observables. Since now $`q_{\mathrm{EA}}`$ and $`m`$ are discontinuous at $`\mathrm{\Gamma }_\mathrm{c}`$ the thermodynamic transition is first order with latent heat and discontinuous susceptibility (see below). The phase diagram resulting from this analysis is represented in Fig. 2 (thin lines). The flat section is the first-order line. We have computed the Edwards-Anderson order parameter and the susceptibility, $`\chi =_0^\beta 𝑑\tau [q_d(\tau )(1m)q_{\mathrm{EA}}]`$, as functions of $`\mathrm{\Gamma }`$ for the $`p=3`$ model. The results are displayed in Fig. 3. The susceptibility has a cusp at $`\mathrm{\Gamma }_\mathrm{c}`$ for $`T>T^{}`$ and a discontinuity for $`T<T^{}`$. The dotted lines correspond to the regions of metastability. Their end points give the amplitude of the ideal hysteresis cycle. The importance of quantum fluctuations may be appreciated from the fact that half way from the transition the order parameter is already reduced by a factor of two. It can be shown analytically that the spectrum of magnetic excitations at $`T=0`$ is gaped both in the PM and SG phases for all $`\mathrm{\Gamma }0`$ (see below, however). Consequently, the latent heat vanishes exponentially as $`T0`$. Since it also vanishes at $`T^{}`$, it must have a maximum at some intermediate temperature. First order quantum transitions were also found in two other models, the fermionic SK-like spin-glass model and a $`p`$-spin model in a transverse field. In contrast, the SG transitions of the Heisenberg EA model and of the SK model in a transverse field are known to be second order. This is also true in finite dimensions . Early experiments on LiHo<sub>x</sub>Y<sub>1-x</sub>F<sub>4</sub> gave some indications that the second order SG transition seen above 25 mK, might become first order at lower temperatures . More recently, hysteresis effects have been observed in this system as a function of the transverse field , giving further support to this idea. The model that we study here captures this phenomenology. We discuss next the consequences of the use of the marginality condition rather than thermodynamics in the determination of the breaking point. In this approach it is not required that $`\mathrm{F}`$ be an extremum with respect to $`m`$ but that the SG phase be marginally stable. This implies that the “replicon” eigenvalue $`\mathrm{\Lambda }`$ must vanish throughout the low-temperature phase. The calculation of $`\mathrm{\Lambda }`$ is analogous to the classical one . The result is $`\mathrm{\Lambda }`$ $`=`$ $`\left[{\displaystyle \frac{\stackrel{~}{q}_d(0)+\beta q_{\mathrm{EA}}(m1)}{\stackrel{~}{q}_d^2(0)\beta ^2q_{\mathrm{EA}}^2(m1)+\stackrel{~}{q}_d(0)\beta q_{\mathrm{EA}}(m2)}}\right]^2\beta ^2`$ (8) $`{\displaystyle \frac{\beta ^2}{2}}p(p1)q_{\mathrm{EA}}^{p2}.`$ The value of $`m`$ follows from the equation $`\mathrm{\Lambda }=0`$, that combined with the equation $`\delta \mathrm{F}/\delta q_{\mathrm{EA}}=0`$, yields $$m=T(p2)\sqrt{2/(p(p1))}q_{\mathrm{EA}}^{p/2}.$$ (9) Notice that this expression is equivalent to Eq. (6) with the substitution $`x_p(p2)`$. Interestingly enough, Eq. (9) is identical to the equation found for the fluctuation-dissipation theorem (FDT) violation parameter, $`X`$, in the real-time dynamical calculation . Moreover, the static and dynamical equations for $`q_{\mathrm{EA}}`$ are also identical which implies that $`m=X`$. The coincidence between the values of $`X`$ and $`m`$ for the marginal SG state has been noticed several times for classical models. This is the first explicit evidence of its validity in a quantum problem. $`X`$ is related to the effective temperature of the system, $`T_{\mathrm{eff}}=X^1T`$, where $`T`$ is the temperature of the thermal bath it is in contact with. Values of $`X1`$ signal the presence of a non-stationary dynamics and of FDT violations. The fact that $`\beta X=\beta m\mathrm{const}`$ when $`T0`$ shows that a non-trivial $`T_{\mathrm{eff}}`$ is generated even when the temperature of the bath vanishes. We have also shown analytically that the internal energy, computed from $`U=(\beta \mathrm{F})/\beta `$ at constant $`m`$, coincides with the long-time limit of the energy per spin as obtained from dynamics . The dynamic transition line $`\mathrm{\Gamma }_\mathrm{d}(T)`$ may be thus identified as the boundary of the region in the $`T\mathrm{\Gamma }`$ plane where the marginally stable SG exists. Below this line, the dynamics of the system becomes non-stationary and FDT violations set in. The dynamic phase diagram for $`p=3`$ is shown in Fig. 2 (thick lines). As in the equilibrium case, $`m`$ is discontinuous across the dashed line. $`\mathrm{\Gamma }_\mathrm{d}`$ lies always above $`\mathrm{\Gamma }_\mathrm{c}`$ suggesting that the equilibrium state can never be reached dynamically starting from an initial state in the PM phase. The two lines are extremely close to each other for $`TT^{}`$. Within the accuracy of our calculations we cannot assert whether they precisely touch at $`T^{}`$, an intriguing possibility. In the region $`T<T^{}`$, $`m`$ varies continuously along $`\mathrm{\Gamma }_\mathrm{d}(T)`$ and vanishes at the quantum critical point. This has a consequence of potential interest for experiment: FDT violations are predicted to appear suddenly rather than gradually as $`\mathrm{\Gamma }_\mathrm{d}`$ is crossed coming from the high $`\mathrm{\Gamma }`$ region for $`T<T^{}`$. The stationary part of the time-dependent susceptibility in the SG phase can be calculated by analytic continuation of $`\stackrel{~}{q}_d^{}(\omega _k)`$. It may be shown that the excitation spectrum of the marginal SG state is gapless . Furthermore, $`\chi ^{\prime \prime }(\omega )`$ may be calculated exactly for $`T,\omega 0`$. The result is $$\underset{\omega 0}{lim}\frac{\chi ^{\prime \prime }(\omega )}{\omega }=\frac{1}{\mathrm{\Gamma }}\left[\frac{2q_{\mathrm{EA}}^{(2p)}}{p(p1)}\right]^{3/4}.$$ (10) A linear excitation spectrum has also been found in the case of the Heisenberg spin glass model . However, a gapless spectrum is not a consequence of Goldstone’s theorem here as our model does not have any continuous symmetry. A more extended discussion of our results will be presented in a forthcoming paper . One of us (CAdSS) acknowledges financial support from the Portuguese Research Council, FCT, under grant BPD/16303/98. LFC and DRG thank the ECOS-Sud program for a travel grant. We thank T. F. Rosenbaum for making available to us the results of Ref. prior to publication. We also thank G. Biroli, J. Kurchan, G. Lozano and M. Rozenberg for useful discussions.
warning/0003/cond-mat0003150.html
ar5iv
text
# Coherent and incoherent dynamic structure functions of the free Fermi gas ## Figure captions * Comparison between the dimensionless Lindhard Function (solid line above) and the dimensionless incoherent and coherent dynamic structure functions (solid and dashed lines below, respectively) at $`\stackrel{~}{q}=0.01`$. * Same as in figure 1 at $`\stackrel{~}{q}=1`$. * Same as in figure 1 at $`\stackrel{~}{q}=1.9`$.
warning/0003/physics0003059.html
ar5iv
text
# Classical “Dressing” of a Free Electron in a Plane Electromagnetic Wave ## I Introduction The behavior of a free electron in a electromagnetic wave is one of the most commonly discussed topics in classical electromagnetism. Yet, several basic issues remain to be clarified. These relate to the question: to what extent can net energy be transferred from an electromagnetic pulse (such as that of a laser) in vacuum to a free electron? These issues are made more complex by quantum considerations, including the role of the “quasimomentum” of an electron that is “dressed” by an electromagnetic wave . As a small step towards understanding of the larger issues, we consider a simpler question here. The response of a free electron to a plane electromagnetic wave is oscillatory motion in the plane perpendicular to the direction of the wave, in the first approximation. Thus, the electron has momentum transverse to the direction of the wave. However, the wave contains momentum only in its direction, and the radiated wave contains no net momentum (in the nonrelativistic limit). How is momentum conserved in this process? The general sense of the answer has been given by Poynting , who noted that an electromagnetic field can be said to contain a flux of energy (energy per unit area per unit time) given by $$𝐒=\frac{c𝐄\times 𝐁}{4\pi },$$ (1) in Gaussian units, where E is the electric field, B is the magnetic field (taken to be in vacuum throughout this paper) and $`c`$ is the speed of light. Poincaré noted that this flow of energy can also be associated with a momentum density given by $$𝐏_{\mathrm{field}}=\frac{𝐒}{c^2}=\frac{𝐄\times 𝐁}{4\pi c},$$ (2) Hence, in the problem of a free electron in a plane electromagnetic wave we are led to seek an electromagnetic field momentum that is equal and opposite to the mechanical momentum of the electron. In this paper we demonstrate that indeed the mechanical momentum of the oscillating electron is balanced by the field momentum in the interference term between the incident wave and the static field of the electron. We are left with some subtleties when we consider the interference between the incident wave and the oscillating field of the electron. ## II Generalities ### A Motion of an Electron in a Plane Wave We consider a plane electromagnetic wave that propagates in the $`+z`$ direction of a rectangular coordinate system. A fairly general form of this wave is $`𝐄_{\mathrm{wave}}`$ $`=`$ $`\widehat{𝐱}E_x\mathrm{cos}(kz\omega t)\widehat{𝐲}E_y\mathrm{sin}(kz\omega t),`$ (3) $`𝐁_{\mathrm{wave}}`$ $`=`$ $`\widehat{𝐱}E_y\mathrm{sin}(kz\omega t)+\widehat{𝐲}E_x\mathrm{cos}(kz\omega t),`$ (4) where $`\omega =kc`$ is the angular frequency of the wave, $`k=2\pi /\lambda `$ is the wave number and $`\widehat{𝐱}`$ is a unit vector in the $`x`$ direction, etc. When either $`E_x`$ or $`E_y`$ is zero we have a linearly polarized wave, while for $`E_x=\pm E_y`$ we have circular polarization. A free electron of mass $`m`$ oscillates in this field such that its average position is at the origin. This simple statement hides the subtlety that our frame of reference is not the lab frame of an electron that is initially at rest but which is overtaken by a wave . If the velocity of the oscillating electron is small, we can ignore the $`𝐯/c\times 𝐁`$ force and take the motion to be entirely in the plane $`z=0`$. Then, (also ignoring radiation damping) the equation of motion of the electron is $$m\ddot{𝐱}=e𝐄_{\mathrm{wave}}(0,t)=e(\widehat{𝐱}E_x\mathrm{cos}\omega t+\widehat{𝐲}E_y\mathrm{sin}\omega t).$$ (5) Using eq. (4) we find the position of the electron to be $$𝐱=\frac{e}{m\omega ^2}(\widehat{𝐱}E_x\mathrm{cos}\omega t+\widehat{𝐲}E_y\mathrm{sin}\omega t),$$ (6) and the mechanical momentum of the electron is $$𝐩_{\mathrm{mech}}=m\dot{𝐱}=\frac{e}{\omega }(\widehat{𝐱}E_x\mathrm{sin}\omega t\widehat{𝐲}E_y\mathrm{cos}\omega t).$$ (7) The root-mean-square (rms) velocity of the electron is $$v_{\mathrm{rms}}=\sqrt{\dot{x}^2+\dot{y}^2}=\frac{e}{m\omega }\sqrt{\frac{E_x^2+E_y^2}{2}}=\frac{eE_{\mathrm{rms}}}{m\omega c}c.$$ (8) The condition that the $`𝐯/c\times 𝐁`$ force be small is then $$\eta \frac{eE_{\mathrm{rms}}}{m\omega c}1,$$ (9) where the dimensionless measure of field strength, $`\eta `$, is a Lorentz invariant. Similarly, the rms departure of the electron from the origin is $$x_{\mathrm{rms}}=\frac{eE_{\mathrm{rms}}}{m\omega ^2}=\frac{\eta \lambda }{2\pi }.$$ (10) Thus, condition (9) also insures that the extent of the motion of the electron is small compared to a wavelength, and so we may use the dipole approximation when considering the fields of the oscillating electron. In the weak-field approximation, we can now use (7) for the velocity to evaluate the second term of the Lorentz force: $$e\frac{𝐯}{c}\times 𝐁=\frac{e^2(E_x^2E_y^2)}{2m\omega c}\widehat{𝐳}\mathrm{sin}2\omega t.$$ (11) This term vanishes for circular polarization, in which case the motion is wholely in the transverse plane. However, for linear polarization the $`𝐯/c\times 𝐁`$ force leads to oscillations along the $`z`$ axis at frequency $`2\omega `$, as first analyzed in general by Landau . For polarization along the $`\widehat{𝐱}`$ axis, the $`x`$-$`z`$ motion has the form of a “figure 8”, which for weak fields ($`\eta 1`$) is described by $$x=\frac{eE_x}{m\omega ^2}\mathrm{cos}\omega t,z=\frac{e^2E_x^2}{8m^2\omega ^3c}\mathrm{sin}2\omega t.$$ (12) If the electron had been at rest before the arrival of the plane wave, then inside the wave it would move with an average drift velocity given by $$v_z=\frac{\eta ^2/2}{1+\eta ^2/2}c,$$ (13) along the direction of the wave vector, as first deduced by McMillan . In the present paper we work in the frame in which the electron has no average velocity along the $`z`$ axis. Therefore, prior to its encounter with the plane wave the electron had been moving in the negative $`z`$ direction with speed given by (13). ### B Field Momentum The fields associated with the electron can be regarded as the superposition of those of an electron at rest at the origin plus those of a dipole consisting of the actual oscillating electron and a positron at rest at the origin. Thus, we can write the electric field of the electron as $`𝐄_{\mathrm{static}}+𝐄_{\mathrm{osc}}`$ and the magnetic field as $`𝐁_{\mathrm{osc}}`$, where the oscillating fields have the pure frequency $`\omega `$ in the low-velocity limit. The entire electromagnetic momentum density can then be written $$𝐏_{\mathrm{field}}=\frac{(𝐄_{\mathrm{wave}}+𝐄_{\mathrm{static}}+𝐄_{\mathrm{osc}})\times (𝐁_{\mathrm{wave}}+𝐁_{\mathrm{osc}})}{4\pi c}.$$ (14) However, in seeking the field momentum that opposes the mechanical momentum of the electron, we should not include either of the self-momenta $`𝐄_{\mathrm{wave}}\times 𝐁_{\mathrm{wave}}`$ or $`(𝐄_{\mathrm{static}}+𝐄_{\mathrm{osc}})\times 𝐁_{\mathrm{osc}}`$. The former is independent of the electron, while the latter can be considered as a part of the mechanical momentum of the electron according to the concept of “renormalization”. We therefore restrict our attention to the interaction field momentum $$𝐏_{\mathrm{int}}=𝐏_{\mathrm{wave},\mathrm{static}}+𝐏_{\mathrm{wave},\mathrm{osc}},$$ (15) where $$𝐏_{\mathrm{wave},\mathrm{static}}=\frac{𝐄_{\mathrm{static}}\times 𝐁_{\mathrm{wave}}}{4\pi c}.$$ (16) and $$𝐏_{\mathrm{wave},\mathrm{osc}}=\frac{𝐄_{\mathrm{wave}}\times 𝐁_{\mathrm{osc}}+𝐄_{\mathrm{osc}}\times 𝐁_{\mathrm{wave}}}{4\pi c}.$$ (17) We recall from eqs. (7) and (12) that the transversemechanical momentum of the oscillating electron has pure frequency $`\omega `$. Since the wave and the oscillating part of the electron’s field each have frequency $`\omega `$, the term $`𝐏_{\mathrm{wave},\mathrm{osc}}`$ contains harmonic functions of $`\omega ^2`$, which can be resolved into a static term plus ones in frequency $`2\omega `$. Hence we should not expect this term to cancel the mechanical momentum. Rather, we look to the term $`𝐏_{\mathrm{wave},\mathrm{static}}`$, since this has pure frequency $`\omega `$. ## III The Momentum $`𝐏_{\mathrm{wave},\mathrm{static}}`$ The static field of the electron at the origin is, in rectangular coordinates, $$𝐄_{\mathrm{static}}=\frac{e}{r^3}(x\widehat{𝐱}+y\widehat{𝐲}+z\widehat{𝐳}),$$ (18) where $`r`$ is the distance from the origin to the point of observation. Combing this with eq. (4) we have $`𝐏_{\mathrm{wave},\mathrm{static}}`$ $`=`$ $`{\displaystyle \frac{e}{4\pi cr^3}}\{\widehat{𝐱}zE_x\mathrm{cos}(kz\omega t)`$ (21) $`+\widehat{𝐲}zE_y\mathrm{sin}(kz\omega t)`$ $`+\widehat{𝐳}[xE_x\mathrm{cos}(kz\omega t)yE_y\mathrm{cos}(kz\omega t)]\}.`$ When we integrate this over all space to find the total field momentum, the terms in $`\widehat{𝐳}`$ vanish as they are odd in either $`x`$ or $`y`$. Likewise, after expanding the cosine and sine of $`kz\omega t`$, the terms proportional to $`z\mathrm{cos}kz`$ vanish on integration. The remaining terms are thus $`𝐩_{\mathrm{wave},\mathrm{static}}`$ $`=`$ $`{\displaystyle _V}𝐏_{\mathrm{wave},\mathrm{static}}`$ (22) $`=`$ $`{\displaystyle \frac{e}{4\pi c}}(\widehat{𝐱}E_x\mathrm{sin}\omega t+\widehat{𝐲}E_y\mathrm{cos}\omega t){\displaystyle _V}{\displaystyle \frac{z\mathrm{sin}kz}{r^3}}`$ (23) $`=`$ $`{\displaystyle \frac{e}{\omega }}(\widehat{𝐱}E_x\mathrm{sin}\omega t+\widehat{𝐲}E_y\mathrm{cos}\omega t)=𝐩_{\mathrm{mech}},`$ (24) after an elementary volume integration. It is noteworthy that the integration is independent of any hypothesis as to the size of a classical electron. Indeed, the integrand of (22) can be expressed as $`\mathrm{cos}\theta \mathrm{sin}(kr\mathrm{cos}\theta )/r^2`$ via the substitution $`z=r\mathrm{cos}\theta `$. Hence, the integral over a spherical shell is independent of $`r`$ for $`kr1`$, and significant contributions to the integral occur for radii up to one wavelength of the electromagnetic wave. This contrasts with the self-momentum density of the electron which is formally divergent; if the integration is cut off at a minimum radius (the classical electron radius), the dominant contribution occurs within twice that radius. Thus, we have demonstrated the principal result of this paper. ## IV The Momentum $`𝐏_{\mathrm{wave},\mathrm{osc}}`$ Several subtleties in the argument appear when we consider the other interference term in the momentum density (15). For this we must first display the electromagnetic fields of an oscillating electron. ### A The Fields $`𝐄_{\mathrm{osc}}`$ and $`𝐁_{\mathrm{osc}}`$ Since we restrict our attention to an electron that oscillates with amplitude much less than a wavelength of the driving wave, and the electron attains velocities that are much less than the speed of light, it is sufficient to use the dipole approximation to the fields of the electron. While these fields are well known, they are typically presented in imaginary notation, of which only the real part has physical significance. This notation is very useful for discussions in which only time-averaged behavior is of interest. However, we wish to consider the details of momentum balance at an arbitrary moment, and it is preferable to use purely real notation. We begin by noting that the retarded vector potential of the oscillating electron at a point r at time $`t`$ can be written $`𝐀_{\mathrm{osc}}(𝐫,t)`$ $`=`$ $`{\displaystyle \frac{e}{c}}{\displaystyle \frac{\dot{𝐱}(t^{}=tr/c)}{r}}`$ (25) $`=`$ $`{\displaystyle \frac{e^2}{m\omega cr}}[\widehat{𝐱}E_x\mathrm{sin}(kr\omega t)+\widehat{𝐲}E_y\mathrm{cos}(kr\omega t)],`$ (26) using eq. (6) for the motion x of the electron. The oscillating part of the scalar potential is obtained by integration of the Lorentz gauge condition: $$𝐀_{\mathrm{osc}}+\frac{1}{c}\frac{\varphi _{\mathrm{osc}}}{t}=0.$$ (27) We find $`\varphi _{\mathrm{osc}}`$ $`=`$ $`{\displaystyle \frac{e^2}{m\omega ^2}}\{E_x[{\displaystyle \frac{kx}{r^2}}\mathrm{sin}(kr\omega t)+{\displaystyle \frac{x}{r^3}}\mathrm{cos}(kr\omega t)]`$ (29) $`+E_y[{\displaystyle \frac{ky}{r^2}}\mathrm{cos}(kr\omega t){\displaystyle \frac{y}{r^3}}\mathrm{sin}(kr\omega t)]\}.`$ The constant static potential is omitted in the above. The scalar potential could also be deduced from the retarded potential of a moving charge. Equation (29) results on expanding the retarded distance to first order in the field strength of the plane wave. The electric and magnetic fields are, of course, found from the potentials via $$𝐁=\times 𝐀\text{and}𝐄=\varphi \frac{1}{c}\frac{𝐀}{t}.$$ (30) The lengthy expressions for the rectangular components of the fields are $`B_{\mathrm{osc},x}`$ $`=`$ $`{\displaystyle \frac{e^2E_y}{m\omega ^2}}\left[{\displaystyle \frac{k^2z}{r^2}}\mathrm{sin}(kr\omega t)+{\displaystyle \frac{kz}{r^3}}\mathrm{cos}(kr\omega t)\right],`$ (31) $`B_{\mathrm{osc},y}`$ $`=`$ $`{\displaystyle \frac{e^2E_x}{m\omega ^2}}\left[{\displaystyle \frac{k^2z}{r^2}}\mathrm{cos}(kr\omega t){\displaystyle \frac{kz}{r^3}}\mathrm{sin}(kr\omega t)\right],`$ (32) $`B_{\mathrm{osc},z}`$ $`=`$ $`{\displaystyle \frac{e^2E_x}{m\omega ^2}}\left[{\displaystyle \frac{k^2y}{r^2}}\mathrm{cos}(kr\omega t){\displaystyle \frac{ky}{r^3}}\mathrm{sin}(kr\omega t)\right]`$ (33) $`+`$ $`{\displaystyle \frac{e^2E_y}{m\omega ^2}}\left[{\displaystyle \frac{k^2x}{r^2}}\mathrm{sin}(kr\omega t)+{\displaystyle \frac{kx}{r^3}}\mathrm{cos}(kr\omega t)\right],`$ (34) and $`E_{\mathrm{osc},x}`$ $`=`$ $`{\displaystyle \frac{e^2E_x}{m\omega ^2}}[({\displaystyle \frac{3kx^2}{r^4}}{\displaystyle \frac{k}{r^2}})\mathrm{sin}(kr\omega t)`$ (38) $`+({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2x^2}{r^3}}+{\displaystyle \frac{3x^2}{r^5}}{\displaystyle \frac{1}{r^3}})\mathrm{cos}(kr\omega t)]`$ $`{\displaystyle \frac{e^2E_y}{m\omega ^2}}[{\displaystyle \frac{3kxy}{r^4}}\mathrm{cos}(kr\omega t)`$ $`+({\displaystyle \frac{k^2xy}{r^3}}{\displaystyle \frac{3xy}{r^5}})\mathrm{sin}(kr\omega t)],`$ $`E_{\mathrm{osc},\mathrm{y}}`$ $`=`$ $`{\displaystyle \frac{e^2E_x}{m\omega ^2}}[{\displaystyle \frac{3kxy}{r^4}}\mathrm{sin}(kr\omega t)`$ (42) $`({\displaystyle \frac{k^2xy}{r^3}}{\displaystyle \frac{3xy}{r^5}})\mathrm{cos}(kr\omega t)]`$ $`{\displaystyle \frac{e^2E_y}{m\omega ^2}}[({\displaystyle \frac{3ky^2}{r^4}}{\displaystyle \frac{k}{r^2}})\mathrm{cos}(kr\omega t)`$ $`({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2y^2}{r^3}}+{\displaystyle \frac{3y^2}{r^5}}{\displaystyle \frac{1}{r^3}})\mathrm{sin}(kr\omega t)],`$ $`E_{\mathrm{osc},\mathrm{z}}`$ $`=`$ $`{\displaystyle \frac{e^2E_x}{m\omega ^2}}[{\displaystyle \frac{3kxz}{r^4}}\mathrm{sin}(kr\omega t)`$ (46) $`({\displaystyle \frac{k^2xz}{r^3}}{\displaystyle \frac{3xz}{r^5}})\mathrm{cos}(kr\omega t)]`$ $`{\displaystyle \frac{e^2E_y}{m\omega ^2}}[{\displaystyle \frac{3kyz}{r^4}}\mathrm{cos}(kr\omega t)`$ $`+({\displaystyle \frac{k^2yz}{r^3}}{\displaystyle \frac{3yz}{r^5}})\mathrm{sin}(kr\omega t)].`$ These expressions can also be deduced from the Liénard-Wiechert forms for the fields of an accelerated charge, keeping terms only to first order in the strength of the plane wave. ### B Components of $`𝐏_{\mathrm{wave},\mathrm{osc}}`$ Since the wave fields have no $`z`$ component, the $`x`$ component of $`𝐏_{\mathrm{wave},\mathrm{osc}}`$ is given by $$P_{\mathrm{wave},\mathrm{osc},x}=\frac{E_{\mathrm{wave},y}B_{\mathrm{osc},z}E_{\mathrm{osc},z}B_{\mathrm{wave},y}}{4\pi c}.$$ (47) From eqs. (33) and (42) we see that both $`B_{\mathrm{osc},z}`$ and $`E_{\mathrm{osc},z}`$ are odd in either $`x`$ or $`y`$. Therefore, the volume integral of $`P_{\mathrm{wave},\mathrm{osc},x}`$ vanishes, and we do not consider it further. Likewise, $`P_{\mathrm{wave},\mathrm{osc},y}`$ vanishes on integration. This confirms the claim made at the end of sec. II that the interference term $`𝐏_{\mathrm{wave},\mathrm{osc}}`$ is not relevant to the balance of transverse momentum between the electron and the fields. However, the $`z`$ component of $`𝐏_{\mathrm{wave},\mathrm{osc}}`$ does not vanish on integration, and requires further discussion. As the details include some surprises (to the author) I present them at length. $`P_{\mathrm{wave},\mathrm{osc},z}=`$ (49) $`{\displaystyle \frac{E_{\mathrm{w},x}B_{\mathrm{o},y}E_{\mathrm{w},y}B_{\mathrm{o},x}+E_{\mathrm{o},x}B_{\mathrm{w},y}E_{\mathrm{o},y}B_{\mathrm{w},x}}{4\pi c}}=`$ $``$ $`{\displaystyle \frac{e^2E_x^2\mathrm{cos}(kzwt)}{4\pi m\omega ^2c}}\left[{\displaystyle \frac{k^2z}{r^2}}\mathrm{cos}(kr\omega t){\displaystyle \frac{kz}{r^3}}\mathrm{sin}(kr\omega t)\right]`$ (50) $``$ $`{\displaystyle \frac{e^2E_y^2\mathrm{sin}(kzwt)}{4\pi m\omega ^2c}}\left[{\displaystyle \frac{k^2z}{r^2}}\mathrm{sin}(kr\omega t)+{\displaystyle \frac{kz}{r^3}}\mathrm{cos}(kr\omega t)\right]`$ (51) $``$ $`{\displaystyle \frac{e^2E_x^2\mathrm{cos}(kzwt)}{4\pi m\omega ^2c}}[({\displaystyle \frac{3kx^2}{r^4}}{\displaystyle \frac{k}{r^2}})\mathrm{sin}(kr\omega t)`$ (53) $`+({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2x^2}{r^3}}+{\displaystyle \frac{3x^2}{r^5}}{\displaystyle \frac{1}{r^3}})\mathrm{cos}(kr\omega t)]`$ $``$ $`{\displaystyle \frac{e^2E_xE_y\mathrm{cos}(kzwt)}{4\pi m\omega ^2c}}[{\displaystyle \frac{3kxy}{r^4}}\mathrm{cos}(kr\omega t)`$ (55) $`+({\displaystyle \frac{k^2xy}{r^3}}{\displaystyle \frac{3xy}{r^5}})\mathrm{sin}(kr\omega t)]`$ $``$ $`{\displaystyle \frac{e^2E_xE_y\mathrm{sin}(kzwt)}{4\pi m\omega ^2c}}[{\displaystyle \frac{3kxy}{r^4}}\mathrm{sin}(kr\omega t)`$ (57) $`+({\displaystyle \frac{k^2xy}{r^3}}{\displaystyle \frac{3xy}{r^5}})\mathrm{cos}(kr\omega t)]`$ $`+`$ $`{\displaystyle \frac{e^2E_y^2\mathrm{sin}(kzwt)}{4\pi m\omega ^2c}}[({\displaystyle \frac{3ky^2}{r^4}}{\displaystyle \frac{k}{r^2}})\mathrm{cos}(kr\omega t)`$ (59) $`({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2y^2}{r^3}}+{\displaystyle \frac{3y^2}{r^5}}{\displaystyle \frac{1}{r^3}})\mathrm{sin}(kr\omega t)].`$ The terms of $`P_{\mathrm{wave},\mathrm{osc},z}`$ that are proportional to $`E_yE_y`$ are odd on both $`x`$ and $`y`$, and so will vanish on integration. We now consider the implications of eq. (55) separately for waves of circular and linear polarization. ### C Circular Polarization For a circularly polarized wave, we have $`E_x^2=E_y^2`$. Consequently the dimensionless measure of field strength is $`\eta =eE_x/m\omega c=eE_y/m\omega c`$, according to (9). The prefactors $`e^2E_x^2/4\pi m\omega ^2c`$ and $`e^2E_y^2/4\pi m\omega ^2c`$ can therefore both be written $`\eta ^2mc/4\pi `$, and have dimensions of momentum. The terms of eq. (55) in $`E_x^2`$ and $`E_y^2`$ can be combined in pairs via the identities $`\mathrm{cos}(kz\omega t)\mathrm{cos}(kr\omega t)+\mathrm{sin}(kz\omega t)\mathrm{sin}(kr\omega t)`$ (60) $`=\mathrm{cos}kz\mathrm{cos}kr+\mathrm{sin}kz\mathrm{sin}kr,`$ (61) and $`\mathrm{sin}(kz\omega t)\mathrm{cos}(kr\omega t)\mathrm{cos}(kz\omega t)\mathrm{sin}(kr\omega t)`$ (62) $`=\mathrm{sin}kz\mathrm{cos}kr\mathrm{cos}kz\mathrm{sin}kr.`$ (63) A detail: the second term of eq. (55) in $`E_x^2`$ contains factors of $`x^2`$, while second term of in $`E_y^2`$ contains factors of $`y^2`$. But during integration, we can replace $`y^2`$ by $`x^2`$, after which the terms can be combined via (61-63). We see already that the volume integral of $`P_{\mathrm{wave},\mathrm{osc},z}`$ will contain no time dependence! On integration, terms such as $`f(x,r)\mathrm{sin}kz`$ and $`g(x,r)z\mathrm{cos}kz`$ that are odd in $`z`$ will vanish. The integrated field momentum is thus, $$p_{\mathrm{wave},\mathrm{osc},z}=_VP_{\mathrm{wave},\mathrm{osc},z}=\frac{\eta ^2mc}{4\pi }I_1=\frac{4}{3}\eta ^2mc,$$ (64) where $`I_1`$ is the volume integral whose integrand is $`{\displaystyle \frac{k^2z}{r^2}}\mathrm{sin}kz\mathrm{sin}kr+{\displaystyle \frac{kz}{r^3}}\mathrm{sin}kz\mathrm{cos}kr`$ (65) $`+\left({\displaystyle \frac{3kx^2}{r^4}}{\displaystyle \frac{k}{r^2}}\right)\mathrm{cos}kz\mathrm{sin}kr`$ (66) $`+\left({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2y^2}{r^3}}+{\displaystyle \frac{3y^2}{r^5}}{\displaystyle \frac{1}{r^3}}\right)\mathrm{cos}kz\mathrm{cos}kr.`$ (67) We return to the significance of eq. (64) after describing the evaluation of integral $`I_1`$. As seen from eq. (64), the integral $`I_1`$ must be dimensionless, although it is apparently a function of the wave number $`k`$. However, the form of (67) indicates that $`I_1`$ is actually independent of the length scale, so we can set $`k=1`$ during integration. To perform the integration we consider a volume element $`r^2drd\mathrm{cos}\theta d\varphi `$ in a spherical coordinate system with angle $`\theta `$ defined relative to the $`z`$ axis. It is more convenient to keep $`z=r\mathrm{cos}\theta `$ as a variable of integration, using $`dz=rd\mathrm{cos}\theta `$. Then the volume integration has the form $$_V=_0^{\mathrm{}}r𝑑r_r^r𝑑z_0^{2\pi }𝑑\varphi .$$ (68) Most terms of (67) are independent of $`\varphi `$, so their $`\varphi `$ integral is just $`2\pi `$. For the terms in $`x^2`$, we have $$_0^{2\pi }x^2𝑑\varphi =r^2\mathrm{sin}^2\theta \mathrm{cos}^2\varphi d\varphi =\pi (r^2z^2).$$ (69) While each of the four main terms of (67) diverges on integration, it turns out that the two terms in $`\mathrm{cos}z`$ taken together are finite (and likewise for the two terms in $`\mathrm{sin}z`$). We find that $$I_1=I_A+I_B=\frac{16\pi }{3},$$ (70) where $`I_A`$ $`=`$ $`2\pi {\displaystyle _0^{\mathrm{}}}𝑑r{\displaystyle \frac{\mathrm{sin}r}{r}}{\displaystyle _r^r}𝑑zz\mathrm{sin}z`$ (72) $`+2\pi {\displaystyle 𝑑r\frac{\mathrm{cos}r}{r^2}𝑑zz\mathrm{sin}z}`$ $`=`$ $`4\pi ,`$ (73) and $`I_B`$ $`=`$ $`\pi {\displaystyle 𝑑r\frac{\mathrm{sin}r}{r}𝑑z\mathrm{cos}z}`$ (77) $`3\pi {\displaystyle 𝑑r\frac{\mathrm{sin}r}{r^3}𝑑zz^2\mathrm{cos}z}`$ $`+\pi {\displaystyle 𝑑r\mathrm{cos}r\left(1+\frac{1}{r^2}\right)𝑑z\mathrm{cos}z}`$ $`+\pi {\displaystyle 𝑑r\frac{\mathrm{cos}r}{r^2}\left(1\frac{3}{r^2}\right)𝑑zz^2\mathrm{cos}z}`$ $`=`$ $`{\displaystyle \frac{4\pi }{3}}.`$ (78) From detailed evaluation of the radial integral, we find that the integrand approaches a constant value as $`r`$ goes to zero, and that the contribution to the integral at large $`r`$ diminishes as $`1/r`$. That is, the principal contribution is from the region $`kr1`$. We are left with the result (64) that the integral of the interference term in the field momentum density has a constant longitudinal term for an electron oscillating in a circularly polarized wave. Recall that we have performed the analysis in a frame in which the electron has no longitudinal momentum. However, as remarked in sec. IIA, prior to its encounter with the wave, the electron had velocity $`v_z=\eta ^2c/2`$ (assuming $`\eta ^21`$), and therefore had initial mechanical momentum $`p_{\mathrm{mech},z}=\eta ^2mc/2`$. So, we would expect that this initial mechanical momentum had been converted to field momentum, if momentum is to be conserved. The result (64) can be described as a kind of “hidden momentum” , whose appearance can be surprising if one ignores the physical processes needed to arrive at the nominal conditions of the problem. We continue to be puzzled as to why the result (64) is 8/3 times larger than that required to satisfy momentum conservation. ### D Linear Polarization Consider now the case of a linearly polarized wave with electric field along the $`x`$ axis. Then $`E_{\mathrm{rms}}=E_x/\sqrt{2}`$, and the prefactors in (55) can be written as $`\eta ^2mc/2\pi `$. The remaining terms in the momentum density $`P_{\mathrm{wave},\mathrm{osc},z}`$ have time dependences that can be expressed as sums of pure frequencies via the identities $`2\mathrm{cos}(kz\omega t)\mathrm{cos}(kr\omega t)`$ (79) $`=`$ $`\mathrm{cos}kz\mathrm{cos}kr+\mathrm{sin}kz\mathrm{sin}kr`$ (82) $`+(\mathrm{cos}kz\mathrm{cos}kr\mathrm{sin}kz\mathrm{sin}kr)\mathrm{cos}2\omega t`$ $`+(\mathrm{cos}kz\mathrm{sin}kr+\mathrm{sin}kz\mathrm{sin}kr)\mathrm{sin}2\omega t,`$ and $`2\mathrm{cos}(kz\omega t)\mathrm{sin}(kr\omega t)`$ (83) $`=`$ $`\mathrm{cos}kz\mathrm{sin}kr\mathrm{sin}kz\mathrm{cos}kr`$ (86) $`+(\mathrm{cos}kz\mathrm{sin}kr+\mathrm{sin}kz\mathrm{cos}kr)\mathrm{cos}2\omega t`$ $`+(\mathrm{sin}kz\mathrm{sin}kr\mathrm{cos}kz\mathrm{cos}kr)\mathrm{sin}2\omega t,`$ Inserting these into eq. (55) and keeping only those terms that are even in $`z`$, we find the integrated field momentum to be $`p_{\mathrm{wave},\mathrm{osc},z}`$ $`=`$ $`{\displaystyle _V}P_{\mathrm{wave},\mathrm{osc},z}`$ (87) $`=`$ $`{\displaystyle \frac{\eta ^2mc}{4\pi }}(I_1+I_2\mathrm{cos}2\omega t+I_3\mathrm{sin}2\omega t),`$ (88) where integral $`I_1=16\pi /3`$ has been discussed in (67-78), $$I_2=I_A+I_B=\frac{8\pi }{3},$$ (89) and integral $`I_3`$ has the integrand, $`{\displaystyle \frac{k^2z}{r^2}}\mathrm{sin}kz\mathrm{sin}kr{\displaystyle \frac{kz}{r^3}}\mathrm{sin}kz\mathrm{cos}kr`$ (90) $`\left({\displaystyle \frac{3kx^2}{r^4}}{\displaystyle \frac{k}{r^2}}\right)\mathrm{cos}kz\mathrm{sin}kr`$ (91) $`+\left({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2y^2}{r^3}}+{\displaystyle \frac{3y^2}{r^5}}{\displaystyle \frac{1}{r^3}}\right)\mathrm{cos}kz\mathrm{cos}kr.`$ (92) On evaluation, $`I_3=0`$. Hence, the longitudinal component of the interference field momentum of a free electron in a linearly polarized wave is $$p_{\mathrm{wave},\mathrm{osc},z}=\frac{4}{3}\eta ^2mc+\frac{2}{3}\eta mc\mathrm{cos}2\omega t.$$ (93) The constant term is the same as that found in eq. (64) for circular polarization, and represents the initial mechanical momentum of the electron that became stored in the electromagnetic field once the electron became immersed in the wave. As for the second term of (93), recall from eq. (12) that for linear polarization the electron oscillates along the $`z`$ axis at frequency $`2\omega `$. Hence the $`z`$ component of the mechanical momentum of the electron is $$p_{\mathrm{mech},z}=m\dot{z}=\frac{\eta ^2mc}{2}\mathrm{cos}2\omega t.$$ (94) The term in $`p_{\mathrm{wave},\mathrm{osc},z}`$ at frequency $`2\omega `$ is $`4/3`$ of the longitudinal component of the mechanical momentum associated with the “figure 8” motion of the electron. Thus, we have not been completely successful in accounting for momentum conservation when the small, oscillatory longitudinal momentum is considered. The factors of 4/3 and 8/3 are presumably not the same as the famous factor of $`4/3`$ that arise in analyses of the electromagnetic energy and momentum of the self fields of an electron . A further appearance of a factor of 8/3 in the present example occurs when we consider the field energy of the interference terms. ## V The Interference Field Energy It is also interesting to examine the electromagnetic field energy of an electron in a plane wave. As for the momentum density (14), we can write $$U_{\mathrm{total}}=\frac{(𝐄_{\mathrm{wave}}+𝐄_{\mathrm{static}}+𝐄_{\mathrm{osc}})^2+(𝐁_{\mathrm{wave}}+𝐁_{\mathrm{osc}})^2}{8\pi },$$ (95) for the field energy density. Again, we no not consider the divergent energies of the self fields, but only the interference terms, $$U_{\mathrm{int}}=U_{\mathrm{wave},\mathrm{static}}+U_{\mathrm{wave},\mathrm{osc}},$$ (96) where $$U_{\mathrm{wave},\mathrm{static}}=\frac{𝐄_{\mathrm{wave}}𝐄_{\mathrm{static}}}{4\pi }.$$ (97) and $$U_{\mathrm{wave},\mathrm{osc}}=\frac{𝐄_{\mathrm{wave}}𝐄_{\mathrm{osc}}+𝐁_{\mathrm{wave}}𝐁_{\mathrm{osc}}}{4\pi }.$$ (98) In general, the interference field energy density is oscillating. Here, we look for terms that are nonzero after averaging over time. We see at once that $$U_{\mathrm{wave},\mathrm{static}}=0,$$ (99) since all terms have time dependence of $`\mathrm{cos}\omega t`$ or $`\mathrm{sin}\omega t`$. In contrast, $`U_{\mathrm{wave},\mathrm{osc}}`$ will be nonzero as its terms are products of sines and cosines: $`U_{\mathrm{wave},\mathrm{osc}}=`$ (100) $``$ $`{\displaystyle \frac{e^2E_x^2\mathrm{cos}(kzwt)}{4\pi m\omega ^2}}[({\displaystyle \frac{3kx^2}{r^4}}{\displaystyle \frac{k}{r^2}})\mathrm{sin}(kr\omega t)`$ (102) $`+({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2x^2}{r^3}}+{\displaystyle \frac{3x^2}{r^5}}{\displaystyle \frac{1}{r^3}})\mathrm{cos}(kr\omega t)]`$ $``$ $`{\displaystyle \frac{e^2E_xE_y\mathrm{cos}(kzwt)}{4\pi m\omega ^2}}[{\displaystyle \frac{3kxy}{r^4}}\mathrm{cos}(kr\omega t)`$ (104) $`+({\displaystyle \frac{k^2xy}{r^3}}{\displaystyle \frac{3xy}{r^5}})\mathrm{sin}(kr\omega t)],`$ $`+`$ $`{\displaystyle \frac{e^2E_xE_y\mathrm{sin}(kzwt)}{4\pi m\omega ^2}}[{\displaystyle \frac{3kxy}{r^4}}\mathrm{sin}(kr\omega t)`$ (106) $`({\displaystyle \frac{k^2xy}{r^3}}{\displaystyle \frac{3xy}{r^5}})\mathrm{cos}(kr\omega t)]`$ $`+`$ $`{\displaystyle \frac{e^2E_y^2\mathrm{sin}(kzwt)}{4\pi m\omega ^2}}[({\displaystyle \frac{3ky^2}{r^4}}{\displaystyle \frac{k}{r^2}})\mathrm{cos}(kr\omega t)`$ (108) $`({\displaystyle \frac{k^2}{r}}{\displaystyle \frac{k^2y^2}{r^3}}+{\displaystyle \frac{3y^2}{r^5}}{\displaystyle \frac{1}{r^3}})\mathrm{sin}(kr\omega t)]`$ $``$ $`{\displaystyle \frac{e^2E_y^2\mathrm{sin}(kzwt)}{4\pi m\omega ^2}}\left[{\displaystyle \frac{k^2z}{r^2}}\mathrm{sin}(kr\omega t)+{\displaystyle \frac{kz}{r^3}}\mathrm{cos}(kr\omega t)\right]`$ (109) $``$ $`{\displaystyle \frac{e^2E_x^2\mathrm{cos}(kzwt)}{4\pi m\omega ^2}}\left[{\displaystyle \frac{k^2z}{r^2}}\mathrm{cos}(kr\omega t){\displaystyle \frac{kz}{r^3}}\mathrm{sin}(kr\omega t)\right].`$ (110) The terms in $`E_xE_y`$ will vanish on integration over volume. The various time averages are $`2\mathrm{cos}(kz`$ $``$ $`\omega t)\mathrm{cos}(kr\omega t)`$ (112) $`=\mathrm{cos}kz\mathrm{cos}kr+\mathrm{sin}kz\mathrm{sin}kr,`$ $`2\mathrm{sin}(kz`$ $``$ $`\omega t)\mathrm{cos}(kr\omega t)`$ (114) $`=\mathrm{sin}kz\mathrm{cos}kr\mathrm{cos}kz\mathrm{sin}kr,`$ $`2\mathrm{cos}(kz`$ $``$ $`\omega t)\mathrm{sin}(kr\omega t)`$ (116) $`=\mathrm{cos}kz\mathrm{sin}kr\mathrm{sin}kz\mathrm{cos}kr,`$ $`2\mathrm{sin}(kz`$ $``$ $`\omega t)\mathrm{sin}(kr\omega t)`$ (118) $`=\mathrm{cos}kz\mathrm{cos}kr+\mathrm{sin}kz\mathrm{sin}kr.`$ After performing the time average on eq. (106), we keep only terms that are even in $`z`$. These terms have the form (67), and so we find that $$u_{\mathrm{int}}=_VU_{\mathrm{wave},\mathrm{osc}}=\frac{e^2(E_x^2+E_y^2)}{8\pi m\omega ^2}I_1=\frac{4}{3}\eta ^2mc^2,$$ (119) for waves of either linear or circular polarization. As with the case of the interference field momentum, this interference field energy is distributed over a volume of order a cubic wavelength around the electron. Being an interference term, its sign can be negative. We can interpret the quantity, $$\frac{u_{\mathrm{int}}}{c^2}=\frac{4}{3}\eta ^2m,$$ (120) as compensation for the relativistic mass increase of the oscillating electron, which scales as $`v_{\mathrm{rms}}^2/c^2`$ and hence as $`\eta ^2`$ (for small $`\eta `$, recall eq. (8)). Indeed, a general result for the motion of an electron in a plane wave of arbitrary strength $`\eta `$ is that its rms relativistic mass, often called its effective mass, is $$m_{\mathrm{eff}}=m\sqrt{1+\eta ^2}.$$ (121) For small $`\eta `$, the increase in mass is $$\mathrm{\Delta }m\frac{1}{2}\eta ^2m.$$ (122) Thus, the decrease in field energy due to the interference terms between the electromagnetic fields of the wave and electron is $`8/3`$ times the mass increase it should compensate. ## VI Discussion ### A Temporary Acceleration We remarked in sec. IIA that the preceding analysis holds in the average rest frame of the electron. If instead the electron had been at rest prior to the arrival of the plane wave, the velocity of the average rest frame would be $`v_z=(\eta ^2/2)/(1+\eta ^2/2)`$. For this, the amplitude of the plane wave is presumed to have a slow rise from zero to a long plateau at strength $`\eta `$, followed by a slow decline back to zero. Once the wave has passed by the electron, the interference field energy, (119), goes to zero since the integral is dominated by the contribution at distances of order a wavelength from the electron. Hence, the electron’s kinetic energy must return to zero (or to its initial value if that was nonzero). A plane wave, or more precisely, a long pulse that is very nearly a plane wave, cannot transfer net energy to an electron. The acceleration of the electron from zero velocity to $`v_z`$ is only temporary, i.e., for the duration of the plane wave pulse. This result was first deduced by di Francia and by Kibble by different arguments. ### B The Radiation Reaction Our analysis of the energy balance of an electron in a plane wave is not quite complete. We have neglected the energy radiated by the electron. Since the rate of radiation is constant (once the electron is inside the plane wave), the total radiated energy grows linearly with time, and eventually becomes large. The interference energy, (119), is constant in time, and hence cannot account for the radiated energy. More to follow….. ## VII Appendix: Liénard-Wiechert Fields As an alternative to the dipole approximation, we consider the use of the Liénard-Wiechert potentials and fields of a moving electron. We have limited our analysis to the case of a weak plane wave ($`\eta `$1), for which the velocity of the electron is always small $`(\beta =v/c1)`$. In this case we may approximate the time-dependent part of the fields of the electron as proportional to the strength of the field of the plane wave (proportional to $`\eta `$. Then we find that the Liénard-Wiechert fields of the electron are the same as the fields in the dipole approximation. We can show this in two ways. First, we verify that the Liénard-Wiechert potentials reduce to eqs. (25) and (29). Second, we can verify directly that the Liénard-Wiechert fields are the same as eqs. (33) and (42). The Liénard-Wiechert potentials are $$\varphi =\left[\frac{e}{R(1\beta \widehat{𝐧}}\right],𝐀=\left[\frac{e\beta }{R(1\beta \widehat{𝐧}}\right],$$ (123) where the electron is at postion x, the observer is at r, their separation is R = r $``$ x, the unit vector $`\widehat{𝐧}`$ is $`𝐑/R`$, and the brackets, \[ \], indicate that quantities within are to be evaluated at the retarded time, $`t^{}=tR/c`$. We work in the average rest frame of the electron. In the weak-field approximation we ignore the longitudinal motion of the electron, (12), which is quadratic in the strength of the plane wave. Then the velocity vector of the electron is $$\beta (t)=\frac{e}{m\omega c}\left(\widehat{𝐱}E_x\mathrm{sin}\omega t\widehat{𝐲}E_y\mathrm{cos}\omega t\right),$$ (124) from eq. (7). The retarded velocity is thus, $`[\beta ]`$ $`=`$ $`\beta (t^{}=tR/c)`$ (125) $`=`$ $`{\displaystyle \frac{e}{m\omega c}}\left(\widehat{𝐱}E_x\mathrm{sin}(kR\omega t)+\widehat{𝐲}E_y\mathrm{cos}(kr\omega t)\right).`$ (126) Distance $`R`$ differs from $`r`$ because the electron’s oscillatory motion takes it away from the origin. However, the amplitude of the motion is proportional to strength of the plane wave. Hence, we may replace $`R`$ by $`r`$ in eq. (125) with error only in the second order of field strength. Since the vector potential includes a factor $`\beta `$ in the numerator, we can replace $`R`$ by $`r`$ and $`1\beta \widehat{𝐧}`$ by 1 in the first order in the field strength of the plane wave. Thus, $$𝐀=\frac{e^2}{m\omega cr}\left(\widehat{𝐱}E_x\mathrm{sin}(kr\omega t)+\widehat{𝐲}E_y\mathrm{cos}(kr\omega t)\right),$$ (127) in agreement with eq. (30). In the scaler potential, we first bring $`\beta `$ to the numerator: $$\varphi \frac{e[1+\beta \widehat{𝐧}]}{[R]}.$$ (128) Unit vector $`[\widehat{𝐧}]`$ differs from unit vector $`\widehat{𝐫}`$ due to the oscillation of the electron, which is proportional to the field strength of the plane wave. For the scalar potential, however, we must expand the factor $`1/[R]`$ to first order in the field strength. Now, $$[R]=|𝐫𝐱(t^{})|=\sqrt{r^22𝐫𝐱(t^{})+𝐱^2(t^{})},$$ (129) with $`𝐱(t^{})`$ $``$ $`{\displaystyle \frac{e}{m\omega ^2}}\left(\widehat{𝐱}E_x\mathrm{cos}\omega t^{}+E_y\mathrm{cos}\omega t^{}\right)`$ (130) $``$ $`{\displaystyle \frac{e}{m\omega ^2}}\left(\widehat{𝐱}E_x\mathrm{cos}\omega (kr\omega t)\widehat{𝐲}E_y\mathrm{cos}(kr\omega t)\right),`$ (131) again approximating $`R`$ by $`r`$ in the arguments of the cosine and sine, accurate to first order in the field strength. Hence, $`{\displaystyle \frac{1}{[R]}}`$ $``$ $`{\displaystyle \frac{1}{r}}(1+𝐫𝐱(𝐭^{}))`$ (132) $``$ $`{\displaystyle \frac{1}{r}}\left\{1e{\displaystyle \frac{\left(xE_x\mathrm{cos}(kr\omega t)yE_y\mathrm{sin}(kr\omega t)\right)}{m\omega ^2r^2}}\right\}.`$ (133) Altogether, $`\varphi `$ $``$ $`{\displaystyle \frac{e}{r}}{\displaystyle \frac{e^2}{m\omega ^2}}\{E_x({\displaystyle \frac{kx}{r^2}}\mathrm{sin}(kr\omega t)+{\displaystyle \frac{x}{r^3}}\mathrm{cos}(kr\omega t))`$ (135) $`+E_y({\displaystyle \frac{ky}{r^2}}\mathrm{cos}(kr\omega t){\displaystyle \frac{y}{r^3}}\mathrm{sin}(kr\omega t))\},`$ in agreement with eq. (29). Similarly, we could proceed from the Liénard-Wiechert fields, $`𝐄`$ $`=`$ $`e\left[{\displaystyle \frac{\widehat{𝐧}\beta }{\gamma ^2(1\beta \widehat{𝐧})^3R^2}}\right]+{\displaystyle \frac{e}{c}}\left[{\displaystyle \frac{\widehat{𝐧}\times \left\{(\widehat{𝐧}\beta )\times \dot{\beta }\right\}}{(1\beta \widehat{𝐧})^3R}}\right],`$ (136) $`𝐁`$ $`=`$ $`[\widehat{𝐧}\times 𝐄].`$ (137) After some work, we find that these fields are the same as eqs. (33-42), to first order in the strength of the plane wave.
warning/0003/hep-ph0003282.html
ar5iv
text
# Non-uniform chiral phase in effective chiral quark models We analyze the phase diagram in effective chiral quark models (the Nambu–Jona-Lasinio model, the $`\sigma `$-model with quarks) and show that at the mean-field level a phase with a periodically-modulated chiral fields separates the usual phases with broken and restored chiral symmetry. A possible signal of such a phase is the production of multipion jets travelling in opposite directions, with individual pions having momenta of the order of several hundred MeV. This signal can be interpreted in terms of disoriented chiral condensates. H. Niewodniczański Institute of Nuclear Physics, PL-31342 Kraków, Poland PACS: 12.39.Fe, 21.65.+f, 12.38.Mh, 64.70.-p The problem of phase transitions in hadronic matter is one of the most important issues in strong-interaction physics. It has been studied theoretically for many years with a variety of methods, with the results gaining importance at the approach of the operation of RHIC. We now believe that in the ($`T`$-$`\mu `$) diagram, where $`T`$ denotes the temperature and $`\mu `$ the baryon chemical potential, one can recognize a quite complicated structure: phases with broken/restored chiral symmetry, confined/deconfined phases, color superconducting phase (in the case of two light flavors) , or color-flavor locked phase (in the case of three light flavors) . In certain limits detailed features of phase transitions are known from first-principle QCD calculations (color superconducting gaps at asymptotically-large baryon densities ), or general symmetry considerations (e.g. the universality class of the chiral phase transition in the massless limit). Lattice calculations provide information for finite-temperature QCD at zero baryon density . However, these “exact” methods are inapplicable to the case of moderate baryon densities (up to a few times the nuclear saturation density), where out of necessity we have to rely on models . In fact, a lot of our experience comes from model calculations, just to mention the expectation that chiral symmetry is eventually restored as the baryon density is increased. In this letter we consider the chiral phase structure for a class of low-energy chiral models: the Nambu –Jona-Lasinio model and the linear $`\sigma `$-model with quarks . These models have been used extensively over the past years to describe low-energy properties of mesons and baryons. They have also been used to describe dense and hot system of the Fermi gas of quarks , in particular to study chiral restoration at high $`T`$ or $`\mu `$. What is usually claimed is that at the mean-field level (synonymous to one-quark-loop or leading-$`N_c`$) the model has two phases: with broken chiral symmetry, i.e. large dynamically-generated quark mass $`M`$, which appears at low $`T`$ and $`\mu `$, and with restored chiral symmetry, with massless (or almost massless) quarks, at higher values of $`T`$ or $`\mu `$. We argue that the situation is much more complicated if one allows for more general mean-field solutions of the model. We show that for a certain range of $`T`$ and $`\mu `$ there appears a non-uniform phase with broken rotational and isospin symmetry. In the $`T`$-$`\mu `$ diagram this phase separates the usual chiral-symmetry-broken phase from the chiral-symmetry-restored phase. The results for the linear $`\sigma `$-model with quarks at $`T=0`$ were obtained a long time ago by Kutschera, Kotlorz, and one of us (WB) . Here we generalize the calculation to finite temperatures, as well as apply it to the Nambu–Jona-Lasinio model. The Lagrangian densities of the $`\sigma `$-model and of the partially-bosonized Nambu–Jona-Lasinio model have the form $`\overline{\psi }[i/g(\sigma +i\gamma _5\tau \pi )]\psi +L_\mathrm{m}(\sigma ,\pi )`$, where $`\psi `$ denotes the quark fields, $`g`$ is the quark-meson coupling constant, and the mesonic part $`L_\mathrm{m}(\sigma ,\pi )`$ is specific to the model. We consider spatially non-uniform, time-independent chiral fields: $$\sigma (\stackrel{}{x})=\frac{M}{g}\mathrm{cos}(\stackrel{}{q}\stackrel{}{x}),\pi _0(\stackrel{}{x})=\frac{M}{g}\mathrm{sin}(\stackrel{}{q}\stackrel{}{x}),\pi _\pm (\stackrel{}{x})=0,$$ (1) where $`M`$ and $`\stackrel{}{q}`$ are parameters, which are determined dynamically. Such an ansatz has been considered for the first time in Refs. in context of the pion condensation in nuclear matter. In the case of vanishing $`\stackrel{}{q}`$ we recover $`\sigma =\frac{M}{g}`$, $`\stackrel{}{\pi }=0`$. The Dirac equation with fields (1) can be solved exactly , which facilitates greatly practical and theoretical applications. The quark spectrum has two branches: $$E_\pm (\stackrel{}{p})=\sqrt{M^2+\stackrel{}{p}^2+\frac{1}{4}\stackrel{}{q}^2\pm \sqrt{M^2\stackrel{}{q}^2+(\stackrel{}{q}\stackrel{}{p})^2}},$$ (2) where $`\stackrel{}{p}`$ labels the momentum of the quark. The basic quantity of our study is the grand thermodynamical potential density $`\mathrm{\Omega }(T,\mu ;M,\stackrel{}{q})`$. Its global minimum with respect to $`M`$ and $`\stackrel{}{q}`$ at fixed $`T`$ and $`\mu `$ determines the ground state of the system subject to conservation of baryon number. We stress that our calculation is self-consistent, i.e. both quark and meson fields are solutions to the equations of motion. At the mean-field (one-quark-loop) level we have explicitly $`\mathrm{\Omega }(T,\mu ;M,\stackrel{}{q})=\mathrm{\Omega }_0(M,\stackrel{}{q})`$ (3) $`2N_cT{\displaystyle \underset{i=\pm }{}}{\displaystyle \frac{d^3p}{(2\pi )^3}\mathrm{ln}\left[\left(1+e^{(\mu E_i(\stackrel{}{p}))/T}\right)\left(1+e^{(\mu +E_i(\stackrel{}{p}))/T}\right)\right]},`$ The medium part of $`\mathrm{\Omega }`$ describes the ideal gas of quarks with the spectrum (2). The vacuum part, $`\mathrm{\Omega }_0`$, depends on the model considered. In the $`\sigma `$-model $$\mathrm{\Omega }_0^\sigma (M,\stackrel{}{q})=\frac{m_\sigma ^2F_\pi ^2}{8}(M^2/M_0^21)^2+\frac{F_\pi ^2}{2}(M^2/M_0^2)\stackrel{}{q}{}_{}{}^{2},$$ (4) where the parameter $`M_0=gF_\pi `$ is the vacuum value of the constituent quark mass, $`F_\pi =93\mathrm{M}\mathrm{e}\mathrm{V}`$ is the pion decay constant, and $`m_\sigma `$ is the vacuum value of $`\sigma `$-meson mass, treated as a model parameter. The first term in (4) comes from the Mexican Hat potential, while the second term is the kinetic energy of the meson fields. In the Nambu–Jona-Lasinio model we have $$\mathrm{\Omega }_0^{\mathrm{NJL}}(M,\stackrel{}{q})=(M^2M_0^2)/(4G)+V(M)+\frac{F_\pi ^2}{2}(M^2/M_0^2)\stackrel{}{q}{}_{}{}^{2}+𝒪(\stackrel{}{q}{}_{}{}^{4}),$$ (5) where $`G`$ is the four-quark coupling constant, and $`V(M)`$ is the Dirac sea contribution to the energy for the case $`\stackrel{}{q}=0`$. Here we adopt the simple 3-momentum regulator, hence $`V(M)=2N_c/\pi ^2_0^\mathrm{\Lambda }𝑑pp^2\left(\sqrt{p^2+M^2}\sqrt{p^2+M_0^2}\right)`$ , where $`\mathrm{\Lambda }`$ is the cut-off parameter. In the calculation presented below we use the parameters of Ref. : $`G=5.01\mathrm{GeV}^2`$, $`\mathrm{\Lambda }=650\mathrm{M}\mathrm{e}\mathrm{V}`$. Equation (5) has been gradient-expanded in powers of $`\stackrel{}{q}`$. The coefficient of the quadratic term is model independent. Higher-order terms depend on the value of $`\mathrm{\Lambda }`$. However, their contribution is small for moderate values of $`\stackrel{}{q}`$ and we can safely drop them. With this simplification the two models become very similar. The same conclusion holds for other variants of the Nambu–Jona-Lasinio model, which differ by the regularization method. In fact, the basic difference between the models is the difference of values of $`\mathrm{\Omega }_0`$ at the minimum and at the point $`M=0`$, $`\stackrel{}{q}=0`$. The larger this difference, the more difficult it is to restore chiral symmetry. The essence of the dynamics of our system can be understood as follows: as we increase the chemical potential, it becomes favorable for the system to develop a non-zero $`\stackrel{}{q}`$. This is because the quark energies on the $`E_{}`$ branch lower with $`\stackrel{}{q}`$, and the energy gain overcomes the repulsion of the meson kinetic term in $`\mathrm{\Omega }_0`$. Thus, for certain $`T`$ and $`\mu `$ the ground state has finite $`\stackrel{}{q}`$. At the further increase of $`\mu `$ the mass of the quark drops to $`0`$, and chiral symmetry is restored. The phase diagram for the Nambu–Jona-Lasinio model with parameters of Ref. is shown in Fig. 1(a). We can see three phases: at low $`\mu `$ and $`T`$ we have the uniform chirally-broken phase, with $`M>0`$ and $`\stackrel{}{q}`$ $`=0`$, wrapped around it is the phase with $`M>0`$ and $`\stackrel{}{q}`$ $`0`$ , and at high $`\mu `$ or $`T`$ we find the chirally-restored phase, with $`M=0`$. The lower line in Fig. 1(a) describes the first-order phase transition between the uniform and non-uniform chirally broken phases. The discontinuities in the $`M`$ and $`\stackrel{}{q}`$ parameters can be read-off from Fig. 1(c). Correspondingly, in the baryon density – temperature diagram (Fig. 1(b)) we notice the region of equilibrium of the two phases (the region between the two lower lines). The upper curve in Fig. 1(a) shows the second-order phase transition between the non-uniform chirally broken phase to the chirally-restored phase. At this phase transition the order parameter $`M`$ vanishes at non-zero $`\stackrel{}{q}`$. Interestingly, the character of the transition between the uniform and non-uniform chiral phases depends on the parameters of the models. In the $`\sigma `$-model for $`m_\sigma =800`$MeV we find a very similar diagram as in Fig. 1. However for $`m_\sigma =1200`$MeV the phase transition is second-order. It changes character around $`m_\sigma =1040`$MeV. There is always the question of applicability and reliability of a model calculation like ours. Firstly, we need to have sufficiently high baryon densities or temperatures to trust the calculation. Physical systems at low densities consist of nucleons, and we need to dissolve them into quarks for the model to be applicable. By geometrical arguments we may expect this to happen at densities of the order of a few nuclear-matter saturation densities. Note that in Fig. 1 this is the region of the non-uniform chirally-broken phase, thus we may hope that the model is valid there. Secondly, the size of the system should be sufficiently large such that our infinite-size calculation is meaningful. It should be larger than $`2\pi /|\stackrel{}{q}|`$ in order to accommodate at least one period of the chiral field. Using the values of $`|\stackrel{}{q}|200600`$ MeV this gives the minimum size in the range $`26`$ fm, easily accessible in heavy-ion collisions. In addition, a more elaborate calculation should incorporate meson fluctuations. At low temperatures pions are easily excited due to their small mass, and such effects should certainly be included. Note, however, that inclusion of meson loops in the study of the quark condensate in Ref. modified, but did not invalidate the mean-field results. Finally, the effect of the explicit breaking of chiral symmetry by the finite current quark mass, $`m`$, should be incorporated. In studies with $`\stackrel{}{q}=0`$ this effect was not quantitatively small. Moreover, with finite $`m`$ chiral restoration is no longer a second-order phase transition, but becomes a smooth cross-over. Another question is whether within the mean-field approximation the solution with fields (1) is the ground state. What we have verified is that for a certain range of $`T`$ and $`\mu `$ we have a lower-energy state than the uniform phase. A priori, there may exist yet lower-energy, up-to-now unknown, mean-field solutions. We conclude with a discussion of a possible signal of the non-uniform chirally broken phase. Suppose that sufficiently large domains of the non-uniform chirally-broken phase are created in the cooling process of the plasma formed in a heavy-ion collision. Such a domain can be treated as a classical source of pions, in full analogy to disoriented chiral condensates (see e.g. ). The coherent decay of such a source would lead to a characteristic signal in the form of two multi-pion “jets”, with pion momenta peaked around $`\pm \stackrel{}{q}`$ (in the rest frame of the source), where $`\stackrel{}{q}`$ is taken at the phase-transition point, and is typically $`200300\mathrm{M}\mathrm{e}\mathrm{V}`$. Indeed, according to Bjorken’s description, pion quanta $`\delta \varphi ^a(x)`$ satisfy the equation $`\left(\mathrm{}+m_\pi ^2\right)\delta \varphi ^a(x)=j^a(\stackrel{}{x})`$, where $`j^a(\stackrel{}{x})`$ is the source, in our case proportional to the classical pion field, i.e. to $`\mathrm{sin}(\stackrel{}{q}\stackrel{}{x})`$. Thus the Fourier transform of $`\delta \varphi ^a`$ picks up the components at $`\pm \stackrel{}{q}`$. Finite size of the domain spreads out the distribution, but if the domain has the size of, say, $`3`$ wavelengths $`2\pi /|\stackrel{}{q}|`$, the half-width of peaks is only $`20\%`$ of $`|\stackrel{}{q}|`$. Of course, larger domains lead to sharper peaks. If multiple domains contribute, then the signal should be folded with the momentum distribution of the domain and averaged over directions. In Eq. (1) we have used the neutral classical pion field. This case leads to production of neutral pion quanta. However, we can rotate ansatz (1) in isospin, which leads to a degenerate solution. Such a configuration will decay into jets of charged pions. In addition, the quanta of the $`\sigma `$ field decay subsequently into pairs on neutral as well as charged pions. A more detailed calculation of the spectrum of pions requires additional assumptions for the dynamics of the phase transitions and is left for a further study. Also, the analysis of the formation and stability of domains is outside of the scope of this paper. We thank Marek Kutschera, Jacek Dziarmaga, and Wojciech Florkowski for many useful conversations. One of us (MS) acknowledges the support of the Polish State Committee for Scientific Research, grant 2P03B 086 14.
warning/0003/astro-ph0003389.html
ar5iv
text
# Modeling Non-Axisymmetric Bow Shocks: Solution Method and Exact Analytic Solutions ## 1 Introduction Supersonic stellar winds shock the surrounding gas and drive expanding bubbles into the interstellar medium. These shocks provide an opportunity to probe the properties of both the driving stellar wind and the ambient medium. If the star is moving with respect to the interstellar gas, the bubble will be distorted into a cometary shape. When the stellar motion is supersonic, we refer to these as stellar wind bow shocks (Baranov, Krasnobaev & Kulikovskii 1971; Dyson 1975). Since the discovery of such bow shocks around young B stars (Van Buren & McCray 1988), bow shocks have been found associated with many classes of objects, such as pulsars (Kulkarni et al. 1992) and cataclysmic variables (e.g. Vela X-1: Kaper et al. 1997); examples include well-known naked-eye stars (e.g. Betelgeuse: Noriega-Crespo et al. 1997). Bow shocks have been proposed as an explanation for cometary, ultracompact HII regions (Van Buren et al. 1990; Mac Low et al. 1991) and as a means of explaining the lifetimes of ultracompact HII regions. In a recent survey of the IRAS database using HiRes processing, Van Buren, Noriega-Crespo, & Dgani (1995) found 58 candidate bow shocks. Non-axisymmetric stellar wind bow shocks occur when a star with an anisotropic wind moves supersonically with respect to the local medium, or if the star has an isotropic wind but moves in an ambient medium containing a transverse density gradient. Models of non-axisymmetric bow shocks are relevant to cometary ultracompact HII regions due to wind-blowing O stars moving supersonically with respect to the surrounding molecular cloud, when the ambient material does not have a constant density. A non-axisymmetric bow shock has also been invoked to explain the morphology of Kepler’s supernova remnant, where the supernova ejecta collide with a non-axisymmetric bow shock generated by the pre-supernova wind (Bandiera 1987; Borkowski, Blondin & Sarazin 1992). Another example is the bow shock due to the head of a jet propagating into a region with a density gradient. Non-axisymmetric, ram-pressure balance models of the collision between a stellar wind and the photoevaporating flow from an externally illuminated circumstellar disk have been given by Henney et al. (1996). A formulation for steady-state non-axisymmetric bow shocks and colliding winds was given by Bandiera (1993). However, Bandiera’s numerical method is sufficiently complicated that a simpler, analytic method is desirable. In this contribution, I present a method for solving the problem of steady-state, momentum-conserving, non-axisymmetric, thin-shell bow shocks and colliding winds. This is an extension of the previous analytic solution method of Wilkin (1996, hereafter Paper I) and of Cantó, Raga, & Wilkin (1996, hereafter Paper II) to non-axisymmetric problems (see also Wilkin 1997a). An outline of the paper is as follows. In § 2, we formulate the problem of the steady-state collision of two winds, and in § 3 we treat the problem of a bow shock resulting from an isotropic wind interacting with a plane-parallel flow containing a transverse density gradient. In § 4, we allow for non-isotropic winds, especially an axisymmetric wind with random orientation of the symmetry axis with respect to the direction of stellar motion. The rate at which kinetic energy is thermalized for the bow shock is treated in § 5. Results and future directions of this research are summarized in § 6. ## 2 Mathematical Formulation and Solution Method ### 2.1 Description of the Collision Surface The hypersonic collision of two winds will in general result in a system of two shocks. A specific example, that of a stellar wind bow shock, is shown schematically in Figure Modeling Non-Axisymmetric Bow Shocks: Solution Method and Exact Analytic Solutions. Because this paper is concerned with steady-state solutions, the colliding winds are assumed to be unchanging in time. The stellar wind bow shock arises when a wind-blowing star moves supersonically with respect to the intersellar gas. In this case, we formulate the problem in the reference frame of the star, so the ambient medium is described as a wind of parallel streamlines impinging on the bow shock. For the collision of two winds in a binary star, we will neglect orbital motion in order to consider a steady-state problem in an inertial reference frame. In steady-state, the amount of mass and momentum within a given volume does not increase, and a flow pattern exists between the two shocks that carries away the mass and momentum deposited by the colliding winds. If the shocks are radiative, post-shock cooling lowers the temperature of the gas and leads to a large compression. We make the strong thin-shell assumption that cooling is so efficient that the shocked shell collapses to an infinitesimally thin layer, with a finite surface density $`\sigma `$ of matter. For this thin-shell assumption to apply, it is also necessary that the post-shock gas not be supported by magnetic fields, which, if well-coupled to the gas, could maintain a finite thickness even in the presence of efficient cooling. We assume there are no other forces such as those due to radiation or magnetic fields. The two shocked winds may in principle be separated by a contact discontinuity. However, there would then be supersonic shear across the interface that is expected to be unstable, leading to a mixing of the two fluids. A detailed treatment of the mixing is beyond the scope of this work. Instead, the mixing is assumed to be instantaneous, so the shell will have a unique velocity $`𝐕_t`$ at any location within it. This velocity represents an average of the turbulent fluctuations that would be present in a more detailed treatment. In steady-state, the geometric shape of the shell is unchanging in time, so the velocity of matter within the infinitely thin layer must be purely tangent to the shell. There will, however, be acceleration normal to the shell, because the fluid typically does not follow a straight path. The assumption that the incident streams are hypersonic, combined with the perfect cooling assumption, means that the flow within the shell will also be hypersonic, and pressure forces may be neglected in describing the motions along the shell.<sup>1</sup><sup>1</sup>1Clearly this assumption must break down for real, finite temperature systems near the stagnation point, because the tangential flow velocity in the shell vanishes at that point. However, pressure forces depend upon the gradient of the pressure, which also vanishes at the stagnation point, so the pressure forces may still be small compared to momentum deposition in the stagnation region (Wilkin 1997b). In any case, beyond a small region near the stagnation point, the tangential flow in the shell will be supersonic, provided the incident flows are supersonic and the shocked fluid cools efficiently. Defining a spherical coordinate system $`(r,\theta ,\varphi )`$, and denoting the radius of the surface by $`r=R(\theta ,\varphi )`$, a complete description of this idealized shell is given by specifying the quantities $`R,\sigma ,𝐕_t`$ as a function of position $`(\theta ,\varphi )`$ within the shell. There are in fact only four independent quantities, because the condition that the motion within the shell be tangential implies that one need only solve for two velocity components. ### 2.2 Previous Treatment of the Problem The conservation laws of mass and momentum may be used to derive the properties of the shell in terms of those of the two incident winds. Since the momentum conservation law has three components, there will be four equations in four unknowns, or including the condition of zero normal velocity, five equations in five unknowns. Because the assumption of vanishing thickness eliminates one spatial variable, these equations will be partial differential equations (PDEs) in two spatial coordinates. Suitable boundary conditions must also be specified. In practice, this means identifying the stagnation (or standoff) point, where the two winds collide head-on, and where for steady-state, the ram pressures of the two fluids balance. The equations may then be integrated away from the standoff point, following the motion of a fluid element. In order to begin the numerical integrations, one generally finds that an expansion about the conditions at the stagnation point is needed. This approach was taken by Bandiera (1993), who derived a set of PDEs in curvilinear coordinates matched to the shape of the shell. These equations were then solved under the assumption of radial, constant velocity, isotropic winds from two point sources. The moving stellar wind bow shock problem is formally obtained by taking the limit of one source placed infinitely far away, while allowing its mass loss rate to be infinite so as to produce a finite density near the other star (at the bow shock). Bandiera noted that for the specific problems cited, the motion within the shell would be along planes. One may readily see this for the bow shock shown in Figure 1. Consider a point on the bow shock surface, and the plane of constant azimuthal angle containing it and the stellar trajectory. The radial wind striking the shell at this point has momentum lying in this plane. Similarly, the ambient medium striking this point has momentum lying in the plane. If the two incident streams mix instantaneously on impact, the resultant momentum must also lie in the same plane. As the fluid flows along the shell, it continues to incorporate momentum contributions lying in the same plane. The collision of radial winds emanating from two point sources gives the same result. Thus, given the perfect mixing asumption, the flow within the shell will lie along planes, a situation we refer to as meridional flow. If the shocked fluid does not mix, the fluid trajectories in the shell are not confined to a plane, and further work is needed to solve this more complicated problem.<sup>2</sup><sup>2</sup>2In the case of axisymmetric problems with shear, one may show that for divergent flow fields such as those encountered in bow shocks and colliding winds, the flow geometry and the fluxes of mass, momentum and angular momentum are the same as for the mixed case (Wilkin, Cantó and Raga 2000). The ensuing geometric simplification allowed Bandiera to construct a two-dimensional grid, consisting of neighboring trajectories within the shell at a set of azimuthal angles. His computational method was to integrate the PDEs for mass, normal momentum, and tangential momentum, by marching along the trajectories within the shell as if the equations were ODEs. By differencing neighboring solutions, numerical values of the cross-streamline derivatives were obtained and supplied to the ODE integrator. Provided the spacing of the two-dimensional grid is sufficiently fine, his method will yield accurate solutions. Bandiera postulated that the dependence of the equations on derivatives of quantities across streamlines was “fictitious,” although he did not succeed in eliminating it from the equations. In this contribution, the solution method will be simplified to an integral approach that avoids the need for PDEs, solving purely algebraic equations. Additionally, the solution may conveniently be obtained in terms of ordinary Cartesian, cylindrical polar, or spherical polar coordinates, and does not require a coordinate system matching the shape of the shell. ### 2.3 The Solution Method The solution method is based upon the observation that thin shells driven by hypersonic winds are momentum-conserving in the vector sense (Paper I). In order to conserve momentum in steady-state, the momentum flux in the shell must be the vector sum of the two incident momentum fluxes, integrated from the standoff point to the point of interest. Such an integration is performed over the area of the shell between two planes of constant azimuthal angle $`\varphi `$, in the limit that their separation $`\mathrm{\Delta }\varphi `$ is infinitesimal (Fig. 1). For axisymmetric flow with no rotational motion about the symmetry axis, no mass or momentum crosses these bounding planes, so the mass and momentum flowing within such a wedge must exactly equal the sum of the mass or momentum fluxes onto the external faces of the bounding surfaces (shocks). Given the known vector momentum flux, one may determine the shape of the shell, since the fluid must move in the direction of its momentum. While this description is complete, the mathematics was subsequently simplified with the inclusion of an additional, angular momentum flux integral (Paper II). In Papers I and II, only constant velocity, isotropic winds were considered. For such winds, the vector momentum flux incident onto the shocked shell is independent of the detailed shape of the shell, so it is possible to specify the flux of momentum onto the shell from each side analytically. The methods of Papers I and II will now be extended to non-axisymmetric bow shocks and colliding winds, provided the flow is meridional. For meridional flow, the bounding surfaces are again planes of constant $`\varphi `$, allowing us to integrate the external fluxes and determine the internal fluxes within the shell. For anisotropic, radial, constant velocity winds, such as those considered by Bandiera, the momentum flux will depend only upon the number of streamlines intersected, and we will be able to obtain solutions analytically. Proceeding with the solution method, we note the location of the stagnation point, where the two streams collide head-on. For the two-wind collision, this point lies between the two stars, along the line connecting their centers. For the bow shock due to a wind-blowing star moving with respect to the ambient medium, it will lie on the stellar trajectory, in the direction of stellar motion. For either problem, the z-coordinate axis is chosen to contain the stagnation point at $`\theta =0`$, with the coordinate origin at the wind source (Fig. 1). In what follows, the wind from the coordinate origin will be referred to as “the wind”, while the “second wind” may be either a radial wind from a second point source, or the ambient medium in the frame of the moving star, where the ambient velocity is $`𝐕_a=V_a\widehat{𝐳}`$. However, we will show explicit formulas only for the bow shock, reserving detailed treatment of the binary star colliding winds for a future paper. One should bear in mind that the methods shown below will be applicable as well for that problem. Returning to the description of a thin slice of the shell bounded by planes of constant $`\varphi `$, the mass, momentum, and angular momentum flux functions per unit azimuthal angle are defined by $$\mathrm{\Phi }_m=\varpi \sigma V_t\mathrm{sec}\eta ,$$ $`(1)`$ $$𝚽=\mathrm{\Phi }_mV_t\widehat{𝐭},$$ $`(2)`$ $$𝚽_J=𝐑\times 𝚽.$$ $`(3)`$ Here $`\varpi =R\mathrm{sin}\theta `$ is the cylindrical radius, and the unit vector at constant $`\varphi `$ tangent to the shell is $`\widehat{𝐭}=\widehat{\varphi }\times \widehat{𝐧}/|\widehat{\varphi }\times \widehat{𝐧}|`$, where $`\widehat{𝐧}`$ is the unit outward normal to the shell. Noting that the arc length element $`ds`$ traced along the shell is given by $`(ds)^2=(dR)^2+R^2(d\theta )^2+R^2\mathrm{sin}^2\theta (d\varphi )^2,`$ the arc length element traced along the shell at constant $`\theta `$ is $`\varpi \mathrm{sec}\eta d\varphi `$, where the angle $`\eta `$ is given by $$\mathrm{tan}\eta =\frac{R}{\varphi }/R\mathrm{sin}\theta .$$ $`(4)`$ For example, the flux of mass crossing a surface of constant polar angle $`\theta `$, between the azimuthal angles $`\varphi `$ and $`\varphi +d\varphi `$, is $`\mathrm{\Phi }_md\varphi `$. If $`V_\theta `$ is the $`\theta `$-component of velocity within the shell, and we write the cylindrical components of the momentum flux function as $`\mathrm{\Phi }_\varpi `$ and $`\mathrm{\Phi }_z`$, then these components are related to the angular momentum flux by $$𝚽_J=\mathrm{\Phi }_mRV_\theta \widehat{\varphi },$$ $`(5)`$ $$𝚽_J=R(\mathrm{\Phi }_\varpi \mathrm{cos}\theta \mathrm{\Phi }_z\mathrm{sin}\theta )\widehat{\varphi }.$$ $`(6)`$ For meridional flow, the angular momentum within the shell will be in the $`\widehat{\varphi }`$-direction, so for the remainder of the discussion we will only need the magnitude of the angular momentum flux $`\mathrm{\Phi }_J=𝚽_J\widehat{\varphi }`$. The mass, momentum, and angular momentum flux functions due to the incident winds are similarly defined. For the first wind, located at the coordinate origin, these are given in terms of the wind density $`\rho _w`$ and velocity $`𝐕_w`$ as $$\mathrm{\Phi }_{m,w}\mathrm{\Delta }\varphi =\rho _w(𝐕_w\widehat{𝐧})𝑑A,$$ $`(7)`$ $$𝚽_w\mathrm{\Delta }\varphi =\rho _w(𝐕_w\widehat{𝐧})𝐕_w𝑑A,$$ $`(8)`$ $$𝚽_{J,w}\mathrm{\Delta }\varphi =\rho _w(𝐕_w\widehat{𝐧})(𝐑\times 𝐕_w)𝑑A.$$ $`(9)`$ The area of integration is that between two planes of constant $`\varphi `$, from the standoff point to polar angle $`\theta `$, following the shape of the collision surface, which is determined along the course of the integrations. The sign of the normal direction is chosen so as to point away from the origin. The flux functions for the ambient medium are strictly analogous, except the unit vector normal to the collision surface would be in the reverse direction: $$\mathrm{\Phi }_{m,a}\mathrm{\Delta }\varphi =\rho _a(𝐕_a\widehat{𝐧})𝑑A,$$ $`(10)`$ $$𝚽_a\mathrm{\Delta }\varphi =\rho _a(𝐕_a\widehat{𝐧})𝐕_a𝑑A,$$ $`(11)`$ $$𝚽_{J,a}\mathrm{\Delta }\varphi =\rho _a(𝐕_a\widehat{𝐧})(𝐑\times 𝐕_a)𝑑A.$$ $`(12)`$ If $`R(\theta ,\varphi )`$ is the radius of the shell, and its partial derivatives are $`R_\theta `$ and $`R_\varphi `$, the surface area element is $$dA=\sqrt{R^2+R_\theta ^2+R_\varphi ^2\mathrm{csc}^2\theta }R\mathrm{sin}\theta d\theta d\varphi .$$ $`(13)`$ The unit outward normal $`\widehat{𝐧}`$ to the surface is given by $$\widehat{𝐧}=\frac{\left(R\widehat{𝐫}R_\theta \widehat{\theta }R_\varphi \mathrm{csc}\theta \widehat{\varphi }\right)}{\sqrt{R^2+R_\theta ^2+R_\varphi ^2\mathrm{csc}^2\theta }}.$$ $`(14)`$ Combining these, $$\widehat{𝐧}dA=\left(R\widehat{𝐫}R_\theta \widehat{\theta }R_\varphi \mathrm{csc}\theta \widehat{\varphi }\right)R\mathrm{sin}\theta d\theta d\varphi .$$ $`(15)`$ For a radial wind, the partial derivatives with respect to $`\theta `$ and $`\varphi `$ do not enter the expression for $`𝐕_w\widehat{𝐧}dA`$, which is simply $$𝐕_w\widehat{𝐧}dA=V_wR^2\mathrm{sin}\theta d\theta d\varphi .$$ $`(16)`$ Similarly, for the ambient medium $`𝐕_a=V_a\widehat{𝐳}=V_a(\widehat{𝐫}\mathrm{cos}\theta \widehat{\theta }\mathrm{sin}\theta )`$, so noting that $`d\varpi /d\theta =R\mathrm{cos}\theta +R_\theta \mathrm{sin}\theta `$, we have $$𝐕_a\widehat{𝐧}dA=V_a\varpi d\varpi d\varphi .$$ $`(17)`$ This eliminates some of the partial derivatives that complicated Bandiera’s very general treatment of the geometry. The rate at which conserved quantities are advected onto the shell does not depend upon the detailed description of the shell, provided we ensure the correct number of streamlines is counted. Using eq.(16), the resulting forms of eqs.(7-9) for the radial stellar wind are $$\mathrm{\Phi }_{m,w}=_0^\theta R^2\rho _wV_w\mathrm{sin}\theta ^{}d\theta ^{},$$ $`(18)`$ $$𝚽_w=_0^\theta R^2\rho _wV_w^2[\widehat{\varpi }\mathrm{sin}\theta ^{}+\widehat{𝐳}\mathrm{cos}\theta ^{}]\mathrm{sin}\theta ^{}d\theta ^{},$$ $`(19)`$ $$\mathrm{\Phi }_{J,w}=0.$$ $`(20)`$ Here the radial unit vector has been written in terms of its cylindrical polar components, and the integrations are performed at constant $`\varphi `$. Because the stellar wind is radial, it imparts no angular momentum to the shell, and the wind angular momentum flux vanishes. Using eq.(17) to simplify eqs.(10-12) for the ambient medium, $$\mathrm{\Phi }_{m,a}=V_a_0^\varpi \rho _a\varpi ^{}𝑑\varpi ^{},$$ $`(21)`$ $$𝚽_a=V_a^2\widehat{𝐳}_0^\varpi \rho _a\varpi ^{}𝑑\varpi ^{},$$ $`(22)`$ $$𝚽_{J,a}=V_a^2\widehat{\varphi }_0^\varpi \rho _a\varpi _{}^{}{}_{}{}^{2}𝑑\varpi ^{}.$$ $`(23)`$ The mass conservation law for the pre-shock radial wind in steady-state is $$\frac{}{r}\left(r^2\rho _wV_w\right)=0.$$ $`(24)`$ Consequently, if the wind is of mass loss rate $`\dot{M}_w`$ and streamline-average speed $`\overline{V}_w`$, its properties are given by $$\rho _wV_w=\frac{\dot{M}_w}{4\pi r^2}f_w(\theta ,\varphi ),$$ $`(25)`$ $$\rho _wV_w^2=\frac{\dot{M}_w\overline{V}_w}{4\pi r^2}g_w(\theta ,\varphi ).$$ $`(26)`$ Here the dimensionless functions $`f_w`$ and $`g_w`$ are normalized to have unit streamline-average over $`4\pi `$ steradians. Note that eq.(26) assumes the stellar wind to be coasting. Given this assumption, $`g_w`$ is independent of the radius. The wind mass and momentum fluxes onto the shell may now be written $$\mathrm{\Phi }_{m,w}=\frac{\dot{M}_w}{4\pi }_0^\theta f_w(\theta ^{},\varphi )\mathrm{sin}\theta ^{}d\theta ^{},$$ $`(27)`$ $$𝚽_w=\frac{\dot{M}_w\overline{V}_w}{4\pi }_0^\theta g_w(\theta ^{},\varphi )[\widehat{\varpi }\mathrm{sin}\theta ^{}+\widehat{𝐳}\mathrm{cos}\theta ^{}]\mathrm{sin}\theta ^{}d\theta ^{}.$$ $`(28)`$ In order to obtain steady-state solutions, the ambient density must be independent of the z-coordinate, although its dependence on the remaining $`\varpi ,\varphi `$ coordinates may be arbitrary. Thus, we assume an ambient density of the form $$\rho _a=\rho _0f_a(\varpi ,\varphi ),$$ $`(29)`$ where $`\rho _0`$ is the value of the ambient density along the stellar trajectory and $`f(0,\varphi )=1`$. The final forms of the ambient flux functions are thus $$\mathrm{\Phi }_{m,a}=\rho _0V_a_0^\varpi f_a\varpi ^{}𝑑\varpi ^{}.$$ $`(30)`$ $$𝚽_a=V_a\widehat{𝐳}\mathrm{\Phi }_{m,a}.$$ $`(31)`$ $$\mathrm{\Phi }_{J,a}=\rho _0V_a^2_0^\varpi f_a\varpi ^2𝑑\varpi ^{}.$$ $`(32)`$ The mass, momentum, and angular momentum flux functions for both the wind and ambient medium are clearly streamline integrals that do not depend upon the detailed shape of the shell - the integration is essentially one over the solid angle of the $`\mathrm{\Delta }\varphi `$ wedge, as seen by the origin. This is a consequence of the fact that the pre-shock media conserve these quantities in steady-state. Suppose we have performed the integrations for the mass, momentum, and angular momentum onto the narrow slice of the shell, for both incident winds, according to eqs.(20,27,28,30-32). In steady-state, these quantities do not accumulate at any location within the shell, but are carried away by the flow within the layer. The conservation laws of mass, momentum, and angular momentum take the form $$\mathrm{\Phi }_m=\mathrm{\Phi }_{m,w}+\mathrm{\Phi }_{m,a},$$ $`(33)`$ $$𝚽=𝚽_w+𝚽_a,$$ $`(34)`$ $$\mathrm{\Phi }_J=\mathrm{\Phi }_{J,w}+\mathrm{\Phi }_{J,a}.$$ $`(35)`$ Now we may describe the properties of the shell explicitly in terms of those of the external winds. Equation (6) yields the shell radius $$R=\frac{\mathrm{\Phi }_J}{\mathrm{\Phi }_\varpi \mathrm{cos}\theta \mathrm{\Phi }_z\mathrm{sin}\theta }.$$ $`(36)`$ This equation combines the formulation of Paper II, which was in terms of azimuthally integrated fluxes, with the original treatment of Paper I, using fluxes in an infinitesimally thin wedge. In eq.(36), each momentum flux function within the shell is to be evaluated as the sum of appropriate source terms, using the conservation laws (eqs.). It is to be stressed that if the flow is not meridional, eq.(5) does not hold, and the solution method is considerably more complicated. However, eqs.(6,36) are still valid provided the angular momentum flux is replaced by its $`\widehat{\varphi }`$ component. From eqs.(1,2), the tangential velocity of matter in the shell is $$V_t\widehat{𝐭}=𝚽/\mathrm{\Phi }_m,$$ $`(37)`$ while the mass surface density in the shell is given by $$\sigma =(\mathrm{\Phi }_m^2/\varpi |𝚽|)\mathrm{cos}\eta .$$ $`(38)`$ ### 2.4 The Applied Torque Method and an Example To provide a specific example of the solution method, as well as a reference solution to be compared to in the following, we consider the simple problem of a bow shock from an isotropic wind in a uniform ambient medium of density $`\rho _a`$. In this case, $`f_w=g_w=f_a=1`$, and the flux functions are $$\mathrm{\Phi }_{m,w}=\frac{\dot{M}_w}{4\pi }(1\mathrm{cos}\theta ),$$ $`(39)`$ $$𝚽_w=\frac{\dot{M}_wV_w}{8\pi }[\widehat{\varpi }(\theta \mathrm{sin}\theta \mathrm{cos}\theta )+\widehat{𝐳}\mathrm{sin}^2\theta ],$$ $`(40)`$ $$\mathrm{\Phi }_{J,w}=0,$$ $`(20)`$ $$\mathrm{\Phi }_{m,a}=\frac{1}{2}\varpi ^2\rho _aV_a,$$ $`(41)`$ $$𝚽_a=\frac{1}{2}\varpi ^2\rho _aV_a^2\widehat{𝐳},$$ $`(42),`$ $$\mathrm{\Phi }_{J,a}=\frac{1}{3}\varpi ^3\rho _aV_a^2.$$ $`(43)`$ Because the stellar wind has uniform speed, we have replaced $`\overline{V}_w`$ with $`V_w`$. One could immediately substitute these results into eq.(36) to determine the shell’s shape. However, before proceeding to apply the formalism, note that the equation for the shape of the shell depends upon a specific combination of three flux functions in eq.(36). One may divide the equation into two parts, one part for each wind, according to $$𝒯_k=\varpi [\mathrm{\Phi }_{\varpi ,k}\mathrm{cot}\theta \mathrm{\Phi }_{z,k}]\mathrm{\Phi }_{\mathrm{J},k},$$ $`(44)`$ and $`k=w`$ or $`a`$ for the wind and ambient (or second wind) sources. Physically, $`𝒯`$ represents the applied torque necessary to compress the fan of streamlines between the stagnation point and position $`R(\theta ,\varphi )`$ into a unidirectional beam possessing the same momentum flux (we shall call $`𝒯`$ the required torque). This is clearly seen by considering the radial wind from the coordinate origin. It has no angular momentum, but an equivalent beam containing the same momentum flux, located at a specific point $`R(\theta ,\varphi )`$ does indeed have angular momentum about the origin, because the beam will not be radial. The required torque for the isotropic wind follows from eqs. (20,40,44): $$𝒯_w=\frac{\dot{M}_wV_w\varpi }{8\pi }(1\theta \mathrm{cot}\theta ).$$ $`(45)`$ Using eqs.(41-44), the torque necessary to compress the ambient streamlines to a unidirectional beam is $$𝒯_a=\frac{1}{6}\varpi ^3\rho _aV_a^2.$$ $`(46)`$ For the ambient medium, the required torque expression simplifies because the ambient medium has no cylindrical radial momentum, and its z-momentum flux is related to its mass flux by eq.(31), so in general, $$𝒯_a=\varpi V_a\mathrm{\Phi }_{m,a}\mathrm{\Phi }_{J,a}.$$ $`(47)`$ Similarly, the radial wind’s required torque is simplified due to its lack of angular momentum about the origin. The solution surface is now defined by $$𝒯_w+𝒯_a=0.$$ $`(48)`$ The interpretation of eq.(48) is that each wind supplies the torque necessary to compress the other wind. These torques are equal and opposite only if the correct value of the radius (or cylindrical radius $`\varpi `$) is used, conserving angular momentum as well as linear momentum. The utility of this approach is that if we vary the properties of one of the winds, holding the other constant, the shape of the shell is affected only by the change in the applied torque from the second wind. Now the applied torque expressions for the wind (eq.) and the ambient medium (eq.), when substituted into eq.(48), quickly recover the solution of Paper I, $$R(\theta )=R_0\mathrm{csc}\theta \sqrt{3(1\theta \mathrm{cot}\theta )},$$ $`(49)`$ where $`R_0`$ is the standoff radius: $$R_0=\sqrt{\frac{\dot{M}_wV_w}{4\pi \rho _aV_a^2}}.$$ $`(50)`$ ## 3 Solutions for Bow Shock with Ambient Density Gradient In this section we assume the asymmetry of the bow shock to be due to a density gradient in the ambient medium, while the the stellar wind is assumed to be isotropic. The next section will treat bow shocks from anisotropic winds. The mass and momentum contributions of the stellar wind to the shell are given by eqs.(39,40,20), and its required torque is given by eq.(45). The ambient medium is assumed to have a plane-parallel density stratification, so in eq.(29), we replace $`f_a(\varpi ,\varphi )`$ simply by $`f_a(x)`$, with $`f_a`$ is normalized so that $`f_a(0)=1`$, and where $$x=\varpi \mathrm{cos}\varphi .$$ $`(51)`$ The integrated mass and angular momentum flux onto the wedge from the ambient medium are $$\mathrm{\Phi }_{m,a}=\rho _0V_a_0^\varpi f_a(\varpi ^{}\mathrm{cos}\varphi )\varpi ^{}𝑑\varpi ^{}.$$ $`(52)`$ $$\mathrm{\Phi }_{J,a}=\rho _0V_a^2_0^\varpi f_a(\varpi ^{}\mathrm{cos}\varphi )\varpi ^2𝑑\varpi ^{}.$$ $`(53)`$ A complete description of the bow shock is now obtained by adding the contributions of both the wind and ambient medium using the conservation laws (eqs.\[33-35\]). ### 3.1 Solution for Exponential Density Law We first obtain the solution for an exponential distribution of ambient density. Let the density scale height be $`H`$. Define for brevity the coordinate $`y=x/H`$, so the density law is $$f_a(x)=\mathrm{exp}(y),$$ $`(54)`$ and the mass and angular momentum flux integrals are $$\mathrm{\Phi }_{m,a}=H^2\rho _0V_a\mathrm{sec}^2\varphi [1\mathrm{exp}(y)(1+y)],$$ $`(55)`$ $$\mathrm{\Phi }_{J,a}=H^3\rho _0V_a^2\mathrm{sec}^3\varphi [2\mathrm{exp}(y)(y^2+2y+2)].$$ $`(56)`$ The ambient momentum flux follows from eq.(31), while the required torque follows from eq.(47), which yields $$𝒯_a=H^3\rho _0V_a^2\mathrm{sec}^3\varphi \left[y2+\mathrm{exp}(y)(y+2)\right].$$ $`(57)`$ Using this torque formula and that for the stellar wind given by eq.(45), the shell’s shape is given by eq.(48), which yields upon simplification $$l^2\mathrm{sec}^2\varphi y^1[y2+\mathrm{exp}(y)(y+2)]=\frac{1}{2}(1\theta \mathrm{cot}\theta ).$$ $`(58)`$ Here $`l=H/R_0`$ is the density scale height in units of the standoff radius. Letting $`y_{ax}`$ be the value of $`y=x/H`$ appropriate for the axisymmetric solution of eq.(49), $$y_{ax}=\frac{R_0}{H}\mathrm{cos}\varphi \sqrt{3(1\theta \mathrm{cot}\theta )},$$ $`(59)`$ eq.(58) takes the simple form $$\frac{1}{y}\left[y2+\mathrm{exp}(y)(y+2)\right]=\frac{1}{6}y_{ax}^2.$$ $`(60)`$ This formula may be solved numerically for $`y(\theta ,\varphi )`$, so letting $`y_s`$ denote the solution for $`y`$ to eq.(60), the shell’s shape is given by $$R(\theta ,\varphi )=H\mathrm{sec}\varphi \mathrm{csc}\theta y_s(\theta ,\varphi ).$$ $`(61)`$ The result is a family of bow shock solutions distinguished by the value of the nondimensional parameter $`l`$. By examining eq.(60) we may deduce several properties of the solution. The left-hand side of the equation is monotonically increasing as a function of $`y`$. The signs of $`y`$ and $`y_{ax}`$ must be the same, since it is due to the $`\mathrm{cos}\varphi `$ factor, so it follows that $`y`$ increases monotonically, although in a very non-linear fashion, as a function of $`y_{ax}`$. Because the parameter $`l`$ does not enter eq.(60), we see that there is one universal solution for the problem, although this solution is scaled by the value of $`R_0`$ and stretched (distorted) depending upon the value of $`H/R_0`$. The solution for the radius at all angles $`\theta ,\varphi `$ and for all values of $`l`$ follows from Fig. Modeling Non-Axisymmetric Bow Shocks: Solution Method and Exact Analytic Solutions. In particular, as $`y\mathrm{}`$, the left hand side of eq.(60) approaches unity, implying a breakout angle for the shell when $`y_{ax}=\sqrt{6}`$, which yields $$1\theta _{\mathrm{}}\mathrm{cot}\theta _{\mathrm{}}=2l^2\mathrm{sec}^2\varphi .$$ $`(62)`$ The opening angle $`\theta _{\mathrm{}}`$ depends upon $`\mathrm{cos}\varphi `$, and applies only to $`\mathrm{cos}\varphi 0`$. In the tail of the bow shock, for $`\mathrm{cos}\varphi <0`$, we have $`y\mathrm{}`$. In this case, $`y`$ depends logarithmically upon $`y_{ax}`$. The most interesting part of the solution, for $`|y|1`$ may be described by an expansion in terms of small $`y_{ax}`$. Noting that for $`\mathrm{cos}\varphi =0`$ the solution is identical to that of eq.(49), we anticipate that an expansion for small $`\mathrm{cos}\varphi `$ corresponds to an expansion in terms of small $`y_{ax}`$ about the standard bow shock solution. Letting $`R_{ax}`$ be the value of $`R(\theta )`$ for that solution (given by eq.), we obtain $$RR_{ax}[1+\frac{1}{4}y_{ax}+\frac{13}{160}y_{ax}^2+\frac{7}{240}y_{ax}^3+\frac{11843}{1075000}y_{ax}^4+\mathrm{}.].$$ $`(63)`$ The behavior of the solution near the stagnation point is given by $$R=R_0\left[1+\frac{\mathrm{cos}\varphi }{4l}\theta +\left(\frac{1}{5}+\frac{13\mathrm{cos}^2\varphi }{160l^2}\right)\theta ^2+\mathrm{}\right].$$ $`(64)`$ Unlike the axisymmetric bow shock, there is a linear term in the behavior near the standoff point, so it is not describable as parabolic with the z coordinate axis as the axis of a parabola, except for the special angles $`\varphi =\pi /2\mathrm{or}3\pi /2`$. As a consistency check, note that as the scale height increases relative to the standoff radius $`(l=H/R_0\mathrm{})`$, the solution reduces to the standard axisymmetric bow shock (Paper I). The lowest order effect near the standoff point is a tilt of the bow shock head, described by the linear term in $`\theta `$. This means that although the wind and ambient streamlines meet head-on at the standoff point, they are not normal to the shell at this location, quite different from the axisymmetric case. This effect is due to the instantaneous mixing we have assumed. If shear is present in the shell, the two colliding flows could have separate stagnation points, which would be located where the incident stream is indeed normal to the shell. These bow shock solutions differ from the standard model (Baranov, Krasnobaev & Kulikovskii 1971) in that the exponential ambient density distribution implies a finite total mass on one side of the bow shock. The solution then more closely resembles those of two-wind collisions (Paper II) where there is a finite opening angle for the bow shock tail. Similar problems include blast waves in finite mass media (Koo & McKee 1990) and wind breakout from a stratified medium (Cantó 1980; Borkowski, Blondin & Harrington 1997). For the high density region, $`\mathrm{cos}\varphi <0`$, the shell’s asymptotic shape in the tail region is $$z\frac{2l^2}{\pi }\varpi \mathrm{sec}^2\varphi \mathrm{exp}\left(\frac{\varpi }{H}\mathrm{cos}\varphi \right).$$ $`(65)`$ The shell continues to be more distorted as it expands due to the $`\mathrm{exp}(y)`$ factor. Because the shell has a finite opening angle on one side, not all wind streamlines intersect the shell, but those with $`\theta >\theta _{\mathrm{}}(\varphi )`$ freely expand to infinity. Further consequences of the finite opening angle of the shell are discussed in §5. ### 3.2 Solution for Linear or Polynomial Density Law Consider a stratification of the ambient medium according to $$f_a(x)=1+a_1x+a_2x^2+\mathrm{}.$$ $`(66)`$ Of course, one must ensure that this expression is non-negative over the domain of interest. The fluxes of mass and angular momentum onto the shell are $$\mathrm{\Phi }_{m,a}=\varpi ^2\rho _0V_a\left[\frac{1}{2}+\frac{a_1}{3}x+\frac{a_2}{4}x^2+\mathrm{}\right],$$ $`(67)`$ $$\mathrm{\Phi }_{J,a}=\varpi ^3\rho _0V_a^2\left[\frac{1}{3}+\frac{a_1}{4}x+\frac{a_2}{5}x^2+\mathrm{}\right].$$ $`(68)`$ The ambient momentum flux follows from eq.(31), while the required torque follows from eq.(47), giving the result $$𝒯_a=\frac{\varpi ^3}{6}\rho _0V_a^2\left[1+\frac{a_1}{2}x+\frac{3}{10}a_2x^2+\mathrm{}+\frac{6(n+1)!}{(n+3)!}a_nx^n+\mathrm{}\right].$$ $`(69)`$ Substitution into eq.(48), we obtain upon simplification $$\frac{\varpi ^2}{R_0^2}\left[1+\frac{a_1}{2}\varpi \mathrm{cos}\varphi +\frac{3a_2}{10}\varpi ^2\mathrm{cos}^2\varphi +\mathrm{}\right]=3(1\theta \mathrm{cot}\theta ).$$ $`(70)`$ Restricting the treatment to a linear density gradient, with only $`a_1`$ non-zero, this equation is cubic in the variable $`\varpi `$. One may solve analytically or numerically for the function $`\varpi (\theta ,\varphi )`$, so the solution surface is then given by $`R(\theta ,\varphi )=\varpi (\theta ,\varphi )\mathrm{csc}\theta `$. Inclusion of higher order terms in the density law simply increases the degree of the polynomial, but the solution is obtained in the same way. The behavior of the solution near the stagnation point is given by $$\frac{R}{R_0}=1\frac{a_1R_0\mathrm{cos}\varphi }{4}\theta +\left(\frac{1}{5}+(\frac{5}{32}a_1^2\frac{3}{20}a_2)R_0^2\mathrm{cos}^2\varphi \right)\theta ^2+\mathrm{}.$$ $`(71)`$ As a consistency check of the solutions, for the case of a uniform ambient medium $`(a_1,a_2,\mathrm{}=0)`$, eqs.(70,71) reduce to the standard solution of Paper I. As a further check, note that this result is consistent with the solution for an exponential mass distribution, if we choose the coefficients corresponding to the exponential distribution of the previous subsection, $`a_1=1/H,a_2=1/2H^2,\mathrm{}`$, recovering eq.(64). ## 4 Treatment of Anisotropic Winds ### 4.1 The Axisymmetric Wind We now relax the assumption that the wind is isotropic to permit an axisymmetric wind, where the axis of symmetry is misaligned with the direction of stellar motion, or in the case of a two-wind collision in a binary star, where the axis does not point to the other star. Treatment of the accelerated wind will be deferred to a future contribution; in this paper the wind is assumed to be coasting. The coordinate axes are chosen so that the z-direction points in the direction of stellar motion, or towards the second star. In terms of a spherical coordinate system, where the azimuthal angle $`\varphi `$ is measured about the z-axis, motion of the shocked fluid in the shell is along planes of constant $`\varphi `$. We also define starred coordinates so that the $`z_{}`$-axis is the symmetry axis of the stellar wind. The two coordinate systems are related by $`\mathrm{sin}\theta _{}\mathrm{cos}\varphi _{}`$ $`=`$ $`\mathrm{sin}\theta \mathrm{cos}\varphi ,`$ (72) $`\mathrm{sin}\theta _{}\mathrm{sin}\varphi _{}`$ $`=`$ $`\mathrm{sin}\theta \mathrm{sin}\varphi \mathrm{cos}\lambda \mathrm{cos}\theta \mathrm{sin}\lambda ,`$ (73) $`\mathrm{cos}\theta _{}`$ $`=`$ $`\mathrm{sin}\theta \mathrm{sin}\varphi \mathrm{sin}\lambda +\mathrm{cos}\theta \mathrm{cos}\lambda .`$ (74) The wind mass and momentum flux densities depend on the polar angle $`\theta _{}`$: $$\rho _wV_w=\frac{\dot{M}_w}{4\pi r^2}f_w(\theta _{}),$$ $`(75)`$ $$\rho _wV_w^2=\frac{\dot{M}_w\overline{V}_w}{4\pi r^2}g_w(\theta _{}).$$ $`(76)`$ The nondimensional functions $`f_w`$ and $`g_w`$ are normalized to have unit average value over $`4\pi `$ steradians. The incident fluxes of mass, momentum, and angular momentum from the wind onto the shell are $`\mathrm{\Phi }_{m,w}(\theta ,\varphi )`$ $`=`$ $`{\displaystyle \frac{\dot{M}_w}{4\pi }}F_w(\theta ,\varphi ),`$ (77) $`𝚽_w(\theta ,\varphi )`$ $`=`$ $`{\displaystyle \frac{\dot{M}_w\overline{V}_w}{4\pi }}𝐆_w(\theta ,\varphi ),`$ (78) $`\mathrm{\Phi }_{\mathrm{J},w}(\theta ,\varphi )`$ $`=`$ $`0.`$ (20) where the nondimensional functions $`F_w`$ and $`𝐆_w=G_{w,\varpi }\widehat{\varpi }+G_{w,z}\widehat{𝐳}`$ are given by $`F_w`$ $`=`$ $`{\displaystyle _0^\theta }f_w\mathrm{sin}\theta ^{}\mathrm{d}\theta ^{},`$ (79) $`𝐆_w`$ $`=`$ $`{\displaystyle _0^\theta }g_w[\widehat{\varpi }\mathrm{sin}\theta ^{}+\widehat{𝐳}\mathrm{cos}\theta ^{}]\mathrm{sin}\theta ^{}\mathrm{d}\theta ^{}.`$ (80) Here $`f_w`$ and $`g_w`$ have argument $`\theta _{}(\theta ^{},\varphi )`$, so the integrations are performed using the transformation eqs.(72-74) to evaluate $`\theta _{}(\theta ,\varphi )`$. We also define a nondimensional function $`T_w`$ associated with the required torque $`𝒯_w`$ to compress the wind streamlines according to $$𝒯_w=\frac{\dot{M}_w\overline{V}_w}{4\pi }\varpi T_w.$$ $`(81)`$ Because the wind’s angular momentum flux vanishes, eq.(44) implies that $$T_w=G_{w,\varpi }\mathrm{cot}\theta G_{w,z}.$$ $`(82)`$ The functions $`f_w(\theta _{}),g_w(\theta _{})`$ may now be expanded in terms of powers of $`\mathrm{cos}\theta _{}`$, $$f_w(\theta _{})=\underset{i=0}{\overset{\mathrm{}}{}}b_i\mathrm{cos}^i\theta _{},$$ $`(83)`$ $$g_w(\theta _{})=\underset{i=0}{\overset{\mathrm{}}{}}c_i\mathrm{cos}^i\theta _{}.$$ $`(84)`$ Defining for brevity $$p=\mathrm{sin}\varphi \mathrm{sin}\lambda ,$$ $`(85)`$ $$q=\mathrm{cos}\lambda ,$$ $`(86)`$ the transformation given by eq.(74) is $`\mathrm{cos}\theta _{}=p\mathrm{sin}\theta +q\mathrm{cos}\theta `$. Denoting the trigonometric integrals by $$\mathrm{I}_{j,k}=_0^\theta \mathrm{sin}^j\theta ^{}\mathrm{cos}^k\theta ^{}d\theta ^{},$$ $`(87)`$ we evaluate $`F_w`$ and $`𝐆_w`$ using the above series, and bringing $`p`$ and $`q`$ outside of the integrals, we finally have $$F_w=\underset{i=0}{\overset{\mathrm{}}{}}b_i\underset{j=0}{\overset{i}{}}p^{ij}q^j\left(\genfrac{}{}{0pt}{}{i}{j}\right)I_{1+ij,j},$$ $`(88)`$ $$𝐆_w=\underset{i=0}{\overset{\mathrm{}}{}}c_i\underset{j=0}{\overset{i}{}}p^{ij}q^j\left(\genfrac{}{}{0pt}{}{i}{j}\right)\left[\widehat{\varpi }I_{2+ij,j}+\widehat{𝐳}I_{1+ij,1+j}\right].$$ $`(89)`$ Here we have used the binomial coefficients to write the sums. The applied torque necessary to compress the wind streamlines to a thin shell, in nondimensional form, is now $$T_w=\underset{i=0}{\overset{\mathrm{}}{}}c_iT_w^{(i)},$$ $`(90)`$ where the responses due to the individual $`cos^i\theta _{}`$ terms are given by $$T_w^{(i)}=\underset{j=0}{\overset{i}{}}p^{ij}q^j\left(\genfrac{}{}{0pt}{}{i}{j}\right)\left[\mathrm{cot}\theta \mathrm{I}_{2+ij,j}\mathrm{I}_{1+ij,1+j}\right].$$ $`(91)`$ ### 4.2 General Solution for Bow Shock Driven by an Axisymmetric, Misaligned Wind The wind description of the previous subsection may now be applied to the problem of a bow shock driven by an anisotropic wind. The wind is assumed to be driven by a star moving at speed $`V_a`$ in a medium of uniform density $`\rho _a`$. We describe the bow shock’s properties in the frame of the star. The standoff radius, defined as the shell radius at $`\theta =0,`$ which corresponds to $`\theta _{}=\lambda `$, is given by $$R_\lambda =\sqrt{\frac{\dot{M}_w\overline{V}_wg_w(\lambda )}{4\pi \rho _aV_a^2}}.$$ $`(92)`$ We also define $`R_0`$ to be the standoff radius for the equivalent isotropic wind $$R_0=\sqrt{\frac{\dot{M}_w\overline{V}_w}{4\pi \rho _aV_a^2}}.$$ $`(93)`$ It is important to recall for the results below that the standoff radius will be $`R_0\sqrt{g_w(\lambda )}`$, where $$g_w(\lambda )=\underset{i=0}{\overset{\mathrm{}}{}}c_i\mathrm{cos}^i\lambda .$$ $`(94)`$ By eqs.(46,48,82), the solution for the shell’s radius is $$R(\theta ,\varphi )=R_0\mathrm{csc}\theta \sqrt{6T_w},$$ $`(95)`$ where $`T_w`$ is given by eqs.(90,91). The total mass and momentum fluxes, including the contribution from the ambient medium, are $$\mathrm{\Phi }_m=\frac{\dot{M}_w}{4\pi }\left[F_w+\frac{1}{2\alpha }\stackrel{~}{\varpi }^2\right],$$ $`(96)`$ $$𝚽=\frac{\dot{M}_w\overline{V}_w}{4\pi }\left[G_{w,\varpi }\widehat{\varpi }+(G_{w,z}\frac{1}{2}\stackrel{~}{\varpi }^2)\widehat{𝐳}\right],$$ $`(97)`$ where $`\stackrel{~}{\varpi }=\varpi /R_0`$, and $`\alpha =V_a/\overline{V}_w`$. To obtain a complete description of the bow shock’s properties, we must specify the velocity of material in the shell and the mass surface density of matter. The velocity is given by the ratio of the mass and momentum fluxes: $$V_t\widehat{𝐭}=V_a\frac{\left[2G_{w,\varpi }\widehat{\varpi }+(2G_{w,z}\stackrel{~}{\varpi }^2)\widehat{𝐳}\right]}{\left[2\alpha F_w+\stackrel{~}{\varpi }^2\right]}.$$ $`(98)`$ The surface density is given by eq.(38), which yields $$\sigma =\rho _aR_0\frac{\left[2\alpha F_w+\stackrel{~}{\varpi }^2\right]^2}{2\stackrel{~}{\varpi }\sqrt{\left[4G_{w,\varpi }^2+(2G_{w,z}\stackrel{~}{\varpi }^2)^2\right]}}\mathrm{cos}\eta .$$ $`(99)`$ In order to evaluate $`\sigma `$, we need to know the angle $`\eta `$. Using eqs.(4,95), we have $$\mathrm{tan}\eta =\frac{1}{2}\mathrm{csc}\theta \frac{1}{T_w}\frac{T_w}{\varphi },$$ $`(100)`$ where differentiation of eqs.(90,91) with respect to $`\varphi `$, which enters only in the variable $`p`$, yields $$\frac{T_w}{\varphi }=\underset{i=1}{\overset{\mathrm{}}{}}c_i\frac{T_w^{(i)}}{\varphi },$$ $`(101)`$ $$\frac{T_w^{(i)}}{\varphi }=\mathrm{cot}\varphi \underset{j=0}{\overset{i1}{}}(ij)p^{ij}q^j\left(\genfrac{}{}{0pt}{}{i}{j}\right)\left[\mathrm{cot}\theta \mathrm{I}_{2+ij,j}\mathrm{I}_{1+ij,1+j}\right].$$ $`(102)`$ Note that the summation limits are restricted so that the terms with $`j=i`$ vanish, including the $`c_0`$ term. In the region of the standoff point, an expansion to second order in $`\theta `$ gives $$RR_\lambda \{1+\theta \frac{pg_w^{}(\lambda )}{4g_w(\lambda )}+\theta ^2[\frac{1}{5}\frac{p^2g_w^2(\lambda )}{32g_w^2(\lambda )}$$ $$+\frac{3(p^2g_w^{\prime \prime }(\lambda )qg_w^{}(\lambda ))}{40g_w(\lambda )}]\}.$$ $`(103)`$ Primes on the function $`g_w`$ represent differentiation with respect to $`q=\mathrm{cos}\lambda `$. As was the case of asymmetry resulting from an ambient density gradient, there is a $`\theta ^1`$ term indicating a tilt of the standoff region relative to the direction of the stellar motion. The mass and momentum flux functions near the standoff point are $$\mathrm{\Phi }_m\frac{\dot{M}_w}{4\pi }\left\{\frac{1}{2}f_w(\lambda )+\frac{1}{2\alpha }g_w(\lambda )\right\}\theta ^2.$$ $`(104)`$ $$𝚽\frac{\dot{M}_w\overline{V}_w}{4\pi }\left\{\frac{1}{3}g_w(\lambda )\widehat{\varpi }+\frac{p}{3}g_w^{}(\lambda )\widehat{𝐳}\right\}\theta ^3.$$ $`(105)`$ Equations (4,28,29) now give the tangential velocity in the shell and the mass per unit area as $$V_t\widehat{𝐭}\frac{2}{3}V_a\theta \frac{g_w(\lambda )\widehat{\varpi }+pg_w^{}(\lambda )\widehat{𝐳}}{\alpha f_w(\lambda )+g_w(\lambda )},$$ $`(106)`$ and $$\sigma \frac{3R_0\rho _a}{4}\frac{(\alpha f_w(\lambda )+g_w(\lambda ))^2}{\sqrt{g_w(\lambda )(g_w^2(\lambda )+p^2(g_w^{}(\lambda ))^2}}.$$ $`(107)`$ One may readily see that these solutions reduce to the results of Paper I for an isotropic wind ($`f_w=g_w=1`$). ### 4.3 Simple Solutions: Quadratic Dependence on $`cos\theta _{}`$ The simplest non-trivial solution is for a linear dependence of mass and momentum fluxes upon $`\mathrm{cos}\theta _{}`$. However, in astrophysical applications one is frequently concerned with stellar winds that have symmetry with respect to the equatorial plane, such as when the asymmetry arises due to rotation of the star. For winds that are symmetric with respect to $`\theta =\pi /2`$, the functions $`f_w(\theta _{}),g_w(\theta _{})`$ will have expansions in terms of even powers of $`\mathrm{cos}\theta _{}`$ only. Thus, we shall give the solution for a wind that depends upon $`\mathrm{cos}^2\theta _{}`$. This is sufficiently general to include as a subset previous models of non-axisymmetric bow shocks (Bandiera 1993; Chen & Huang 1997). We assume the mass and momentum fluxes are described by $$f_w=b_0+b_1\mathrm{cos}\theta _{}+b_2\mathrm{cos}^2\theta _{},$$ $`(108)`$ $$g_w=c_0+c_1\mathrm{cos}\theta _{}+c_2\mathrm{cos}^2\theta _{},$$ $`(109)`$ where the normalization requires $`b_0=(1b_2/3)`$ and $`c_0=(1c_2/3)`$. The mass flux integral $`F_w`$ is $$F_w=\{b_0^{}(1\mu )+\frac{b_1}{2}[p(\theta \mathrm{sin}\theta \mathrm{cos}\theta )+q\mathrm{sin}^2\theta ]$$ $$+\frac{b_2}{3}\mathrm{sin}^2\theta [(q^2p^2)\mathrm{cos}\theta +2pq\mathrm{sin}\theta ]\},$$ $`(110)`$ where $`b_0^{}=b_0+b_2(2p^2+q^2)/3`$. The components of $`𝐆_w`$ are $$G_{w,\varpi }=\frac{1}{4}\{c_0^{}(2\theta \mathrm{sin}2\theta )+\frac{c_1}{3}[p(89\mathrm{cos}\theta +\mathrm{cos}3\theta )+4q\mathrm{sin}^3\theta ]$$ $$+c_2\mathrm{sin}^3\theta [2pq\mathrm{sin}\theta +(q^2p^2)\mathrm{cos}\theta ]\},$$ $`(111)`$ $$G_{w,z}=\frac{1}{4}\{c_0^{}(1\mathrm{cos}2\theta )+\frac{c_1}{3}[4p\mathrm{sin}^3\theta +q(43\mathrm{cos}\theta \mathrm{cos}3\theta )]$$ $$+c_2[\frac{pq}{4}(4\theta \mathrm{sin}4\theta )+\frac{(q^2p^2)}{2}(2+\mathrm{cos}2\theta )\mathrm{sin}^2\theta ]\}.$$ $`(112)`$ Here the coefficient $`c_0^{}=c_0+c_2(3p^2+q^2)/4`$ depends on $`\varphi `$ and $`\lambda `$. Using eqs.(82,111,112) we obtain $$T_w=\{\frac{c_0^{}}{2}(1\theta \mathrm{cot}\theta )\frac{c_1}{3}(1\mathrm{cos}\theta )[q+p\mathrm{tan}(\frac{\theta }{2})]$$ $$+\frac{c_2}{8}[(p^2q^2)\mathrm{sin}^2\theta pq(2\theta \mathrm{sin}2\theta )]\}.$$ $`(113)`$ Using eq.(95), with $`T_w`$ given by eq.(113), we obtain the shell’s shape $$R=R_0\mathrm{csc}\theta \{3(1\theta \mathrm{cot}\theta )(c_0+\frac{c_2}{4}(3p^2+q^2))$$ $$+2c_1(1\mathrm{cos}\theta )[q+p\mathrm{tan}(\frac{\theta }{2}\left)\right]$$ $$+\frac{3c_2}{4}[(q^2p^2)\mathrm{sin}^2\theta +pq(2\theta \mathrm{sin}2\theta )]\}^{1/2},$$ $`(114)`$ where the explicit dependence upon stellar inclination and azimuthal angle is obtained using $`p=\mathrm{sin}\varphi \mathrm{sin}\lambda `$ and $`q=\mathrm{cos}\lambda `$. Examples are shown in Figures 4-6 for even parity winds with $`c_2=3,1.5`$, and $`1.5`$. Noting that Bandiera’s notation for this problem was $`\mathrm{\Delta }_Y=2c_2/3`$, these correspond to his cases of $`\mathrm{\Delta }_Y=2,1`$ and $`+1`$. In the neighborhood of the standoff point, the shell’s shape is given by $$R^2/R_0^2[(c_0+c_1q+c_2q^2)+\frac{p}{2}(c_1+2c_2q)\theta $$ $$+(8c_0+5c_1q+2c_2(3p^2+q^2))\theta ^2/20].$$ $`(115)`$ ### 4.4 The Non-Axisymmetric Wind The previous formulation is not restricted to an axisymmetric wind, so one may obtain solutions for bow shocks driven by radial, non-axisymmetric winds, by defining the momentum flux from the wind in terms of spherical harmonics. Because the method is clear from the previous discussion of an axisymmetric wind, we will not give further details here. We note only that the response from each spherical harmonic contribution to the wind momentum flux may be obtained from the required torque formula. Then the shape of the shell will depend upon the square root of the sum of these responses. ## 5 Energy Thermalization Rates The energy thermalization rate of axisymmetric bow shocks and colliding winds has been considered by Wilkin, Cantó & Raga (2000). The thermalization occurs in the shocks, post-shock (radiative) relaxation, and mixing. The maximum possible rate of thermalization is given by the total incident kinetic energy flux in the center of mass frame. Here we are concerned with the total rate of energy thermalized, over the entire shell, rather than the amount per unit area. By center of mass, we mean the center of mass of the matter deposited onto the shell per unit time. This maximum thermalization rate is for the case of complete mixing. For the bow shock in a uniform ambient medium, the center of mass frame corresponds to the ambient frame, because a formally infinite amount of ambient mass per unit time strikes the shell, while the stellar wind has finite mass loss rate. For the bow shock driven by a star with a radial wind, the total energy thermalization rate is given by $$\dot{E}_{\mathrm{therm}}=_0^{2\pi }_0^\pi \frac{1}{2}R^2\rho _wV_w^2(𝐕_w^{}𝐕_{\mathrm{shell}}^{})\widehat{𝐫}\mathrm{sin}\theta d\theta d\varphi .$$ $`(116)`$ Here primes indicate velocities in the ambient frame, so $`𝐕_w^{}=V_w\widehat{𝐫}+V_a\widehat{𝐳}`$, giving $$V_w^2=[V_w^2+2V_wV_a\mathrm{cos}\theta +V_{}^2].$$ $`(117)`$ The velocity of the shell is its pattern speed and corresponds to the stellar velocity $`𝐕_{}=V_a\widehat{𝐳}`$. Because we are assuming a non-accelerating wind, the integral depends only upon the streamlines and not the detailed shape of the shell, so we may perform the integration over a spherical surface. For the case of an isotropic wind from a star moving through a plane-parallel stratified ambient medium, the rate of energy thermalization is precisely the same as for a uniform medium (Wilkin, Cantó, & Raga 1997), since it is the kinetic energy in the ambient frame that is thermalized: $$\dot{E}_{\mathrm{therm}}=\frac{1}{2}\dot{M}_w(V_w^2+V_{}^2).$$ $`(118)`$ This rate of energy thermalization is appropriate for an arbitrary ambient distribution of matter, provided only that it is truly infinite in mass, and that the bow shock is steady-state in the star’s frame. The second requirement demands that the ambient mass distribution be independent of the z-coordinate, although its $`(\varpi ,\varphi )`$ distribution is arbitrary. An exception, for example, is the case of an exponential stratification, for which case the total ambient mass flux may be finite (on one side of the bow shock). For this case, the center of mass frame is not equal to the ambient frame, and it is important to separately define the center of mass frame for each $`\varphi `$ slice, as it will depend upon azimuthal angle for the portion of the bow shock that contains finite opening angle. Because of the resulting opening angle, not all of the stellar wind kinetic energy (in the ambient frame) will be thermalized, for two reasons: (1) some streamlines miss the shell, and (2) there is a net flow of momentum along the shell even at infinite distance, whereas for the standard bow shock $`V_t0`$ in the ambient frame, far in the bow shock tail. We now consider the case of an anisotropic wind, to determine whether the energy thermalization rate depends on the orientation of the wind. However, the thermalization rate remains independent of the stratification of the ambient medium, subject to the caveat about finite-mass systems. For the remainder of this section, we allow the wind to be non-axisymmetric, so in equations (75,76) we use $`f_w=f_w(\theta _{},\varphi _{}),g_w=g_w(\theta _{},\varphi _{})`$. When we refer to an axisymmetric wind, its axis of symmetry will be $`\theta _{}=0`$ as before. The total kinetic energy loss rate of the wind is $$\dot{E}_w=\frac{1}{8\pi }\dot{M}_w\overline{V}_w^2_0^{2\pi }_0^\pi \frac{g_w^2}{f_w}\mathrm{sin}\theta _{}d\theta _{}d\varphi _{}.$$ $`(119)`$ The functions $`f_w`$ and $`g_w`$ are assumed to have unit average over $`4\pi `$ steradians. In terms of this energy loss rate, we define the mean square wind speed as $$<V_w^2>=\mathrm{\hspace{0.17em}2}\dot{E}_w/\dot{M}_w.$$ $`(120)`$ We also define a total vector momentum flux for the wind, a quantity that is non-vanishing only if the wind is not point-symmetric with respect to the origin $$𝐏_w=\frac{\dot{M}_w\overline{V}_w}{4\pi }_0^{2\pi }_0^\pi g_w\widehat{r}\mathrm{sin}\theta d\theta d\varphi ,$$ $`(121)`$ $$=_0^{2\pi }𝚽_w(\pi ,\varphi )𝑑\varphi .$$ $`(122)`$ We now wish to calculate the kinetic energy deposition rate to the shell in the ambient frame. Letting $`\alpha =V_{}/\overline{V}_w`$, the wind velocity (squared) in the ambient frame becomes $$V_w^2=\overline{V}_w^2\left[\frac{g_w^2}{f_w^2}+2\alpha \frac{g_w}{f_w}\mathrm{cos}\theta +\alpha ^2\right].$$ $`(123)`$ The wind density is given by $$4\pi R^2\rho _w=\frac{\dot{M}_w}{\overline{V}_w}\frac{f_w^2}{g_w}.$$ $`(124)`$ The total kinetic energy flux onto the shell in the ambient frame is $$\dot{E}_{\mathrm{therm}}=\frac{\dot{M}_w\overline{V}_w^2}{8\pi }_0^{2\pi }_0^\pi \left[\frac{g_w^2}{f_w}+2\alpha g_w\mathrm{cos}\theta +\alpha ^2f_w\right]\mathrm{sin}\theta _{}d\theta _{}d\varphi _{}.$$ $`(125)`$ The normalization condition for $`f`$ permits the last term to be easily integrated. The second term is $`V_{}`$ times the $`z`$ component of the wind momentum flux. In terms of the total wind kinetic energy flux and momentum flux, we have $$\dot{E}_{therm}=\frac{1}{2}\dot{M}_w(<V_w^2>+V_{}^2)+𝐏_w𝐕_{}.$$ $`(126)`$ We see that a sufficient condition for the total thermalization rate to be independent of the wind orientation is the vanishing of $`P_{z,w}`$. Because we do not wish to invoke a special alignment of the wind orientation with the star’s direction of motion, more typically this requires a point-symmetric wind such that $`𝐏_w=0`$. As a special case, this can be more explicitly confirmed for the axisymmetric wind, assuming it is symmetric with respect to its midplane $`\theta _{}=\pi /2`$. Because we are considering the total thermalization rate, we may perform the solid angle integration in starred coordinates. In this case, we need only the transformation equation for $`\mathrm{cos}\theta `$ which is given by eq.(74). We assume the wind to be axisymmetric and symmetric with respect to $`\theta _{}=\pi /2`$, so it has an expansion in $`\mathrm{cos}\theta _{}`$ with only even powers. This implies that the $`2\alpha \mathrm{cos}\theta `$ term in the integral will give no contribution, by parity, because it yields only a $`\mathrm{sin}\varphi _{}`$ term, which vanishes in the azimuthal integration and terms with odd powers of $`\mathrm{cos}\theta _{}`$, which vanish in the polar angle integration. The remaining contribution is then $$\dot{E}_{\mathrm{therm}}=\frac{\dot{M}_w\overline{V}_w}{4}_0^\pi \left[\frac{g_w^2(\theta _{})}{f_w(\theta _{})}+\alpha ^2f_w(\theta _{})\right]\mathrm{sin}\theta _{}d\theta _{},$$ $`(127)`$ $$=\frac{1}{2}\dot{M}_w(<V_w^2>+V_{}^2).$$ $`(128)`$ The total energy thermalization rate is independent of the orientation of the midplane-symmetric, axisymmetric wind. For the case of a non-point-symmetric wind, including an axisymmetric wind that doesn’t possess mirror symmetry with respect to the equator, the total thermalization rate will depend upon the stellar orientation because of the contribution from the $`2\alpha \mathrm{cos}\theta `$ term. This is the case of the example solution given in Section 4.3, for the wind with quadratic dependence on $`\mathrm{cos}\theta _{}`$. For that wind, the total vector momentum loss rate is $$𝐏_w=\frac{\dot{M}_w\overline{V}_w}{3}c_1\widehat{𝐳}_{},$$ $`(129)`$ where $`\widehat{𝐳}_{}`$ points in the direction of the wind symmetry axis $`z_{}`$. The thermalization rate for the bow shock driven by this wind therefor has a fluctuation depending on orientation of the wind symmetry axis with respect to the stellar velocity of magnitude $$\mathrm{\Delta }\dot{E}_{therm}=\frac{\dot{M}_w\overline{V}_wV_{}}{3}c_1\mathrm{cos}\lambda .$$ $`(130)`$ If the quadratic term in the wind momentum vanishes ($`c_2=0`$), then the absolute value of $`c_1`$ may not exceed unity. The amount this changes the total thermalization rate then depends upon the parameter $`\alpha =V_{}/V_w`$. We must note that changing orientation of the point-symmetric wind did not change the total thermalization rate, but the spectrum of shock emission will clearly be different. For example the peak post-shock temperature will depend upon the incident normal component of velocity. The sum of thermalization by shock, relaxation, and mixing is independent of orientation, but the individual contributions will vary. This highlights the fact that the properties of bow shocks due to non-axisymmetric winds may be determined by measuring the bow shock shape, mass surface density (or column density), kinematics, and radiated energy. Although detailed shock calculations are beyond the scope of this work, a substantial amount of information should be obtainable from these global quantities and may be sufficient for sources where the data are sparse. ## 6 Summary I have shown how to solve the problem of non-axisymmetric bow shocks and wind collisions with a simple formalism that requires only algebraic equations. Most often one considers constant wind speed for such problems, and in this case the method leads to exact, analytic (although possibly implicit) solutions. The availablility of simple analytic solutions makes it much easier to model observed sources and derive the properties of the driving winds. Among the principal applications of these solutions would be to determine the cause of the asymmetry in observed bow shocks, whether it be due to the ambient medium or an anisotropic wind, or both. Future improvements are needed to include non-radial and accelerating winds and shearing motions in the shell, in which case the fluid elements are not restricted to a plane. Also, non-axisymmetric bow shocks due to colliding winds in binary systems, including the orbital motion, require a formulation in a non-inertial frame. A forthcoming paper will describe how to solve the general problem of colliding winds from two stars, including both anisotropy and acceleration in the winds. Portions of this work were done as an NRC Research Associate at NASA Ames Research Center, and at IPAC with support by Long Term Space Astrophysics grant to D.Van Buren. IPAC is operated by JPL and the California Institute of Technology under contract with the National Aeronautics and Space Administration. I am greatful for helpful discussions with J.Anderson and C.McKee.
warning/0003/cond-mat0003092.html
ar5iv
text
# 1 Introduction ## 1 Introduction Fragmentation is an irreversible kinetic phenomenon which occurs in many physical and chemical processes. Because of broad range of applications, many recent studies have been carried out to investigate the kinetics of the processes by introducing simple fragmentation models . In Ziff and MaGrady presented a model of fragmentation in which the rate of break up depends on the size of the fragments. A general discussion of the kinetics of continuous fragmentation processes was given by Cheng and Redner in . In Krapivsky and Ben-Naim studied the kinetics of random fragmentation of multidimensional objects. In three models of binary fragmentation were investigated analytically in which at each time step, the largest fragment in the system is broken with some externally tuneable probability. In the second model of that paper, at each time step, there is a fixed rate of fragmentation for the largest fragment and a rate proportional to the inverse of the fragment size for all other fragments smaller than the largest one. The model was solved exactly in the long time limit to reveal a power law distribution with an exponent, which depends on the precise details of the fragmentation process. Surprisingly, there is a similarity between the result of the stable distribution from the second model of , with some experimental and analytical results of fragment size and mass distributions with power law forms in the shock fragmentation . In a generalization of the second model of was carried out. Some models of binary fragmentation were introduced which revealed composite power law distributions for the mass and size of the fragments . Another behaviour of distribution which is observed in many different processes has a log-normal form . In a comprehensive study in the log-normal distribution is given. The application of these distributions in many different areas of science is discussed in , such as economics, biology, astronomy, philology, small particle statistics and physical and industrial processes. The log-normal behaviour is known to be able to describe the size distribution well in a wide variety of geological situations , such as that of rocks in a boulder field. In an analytical and numerical study was carried out on the fragmentation of long thin glass rods. Some stochastic models for one-dimensional brittle fracture were introduced. The models successively extended to describe cascade fracture of long thin glass rods. They are found to give the fragment size distribution close to a log-normal one. In terms of rupture in liquids, experiments in rupture of mercury drops were performed . The result of size distribution for the cumulative number of droplets showed a log-normal distribution at the small falling heights. A transition of the distribution from log-normal to scaling behaviour was observed as the falling height was increased. In an experimental work on the shock fragmentation of long thin glass rods was carried out by Ishii and Matsushita. The results of fragment size and mass distributions at small falling heights showed a log-normal form for larger fragments and a power law form for smaller fragments. The crossover was seen to be at length scales around the rod diameter. The mass and size distributions for the larger fragments showed a power law form as the falling height was increased. This paper is organised as follows. In sections 2 and 3 we introduce two new models, which are generalisations of the second model of and the models of fragmentation in . In the models we introduce a time dependent transition size which produces two regions of fragment sizes with different rates of fragmentation above and below the transition size. A rate of fragmentation proportional to the inverse of the fragment size in the smaller size region produces a power law distribution in that region. A rate of fragmentation combined of two terms, one proportional to the inverse of the fragment size and the other proportional to a logarithmic function of the fragment size yield a log-normal distribution in the larger size region. Finally in section 4 we summarise our conclusions and discuss the application of our results to experimental work on the shock fragmentation of long thin glass rods and the rupture of mercury drops . For the rest of this introduction we give a very brief summary of previous work on models of binary fragmentation. In these models, the density of fragments of size $`y`$ at time $`t`$, $`n(y,t)`$, evolves according to $$n(y,t+\delta t)=n(y,t)\delta tn(y,t)_0^yR(z,yz)𝑑z+\mathrm{\hspace{0.17em}2}\delta t_y^{L(t)}R(y,zy)n(z,t)𝑑z,$$ (1) where $`R(y,z)`$ is the intrinsic rate that a fragment of size $`(y+z)`$ breaks into fragments of size $`y`$ and $`z`$. The first term on the right hand side is the contribution from those fragments not chosen for the fragmentation in time $`(t,t+\delta t)`$. The second term represents the decrease in the number of fragments of size $`y`$ by the fragmentation into fragments of size $`z`$ and $`(yz)(y>z)`$. The third term represents the increase in the number of particles of size $`y`$ due to the fragmentation of particles of size $`z(>y)`$, such that one of the products is of size $`y`$. The upper limit of the second integral on the right-hand side is usually set to some fixed value greater than the size of the largest particle in the system. This value is usually $`1`$ or $`\mathrm{}`$ depending on the details of the model. To connect with the analysis of our models in the following sections, we have set the upper limit to $`L(t)`$, the size of the largest particle at time $`t`$. As $`n(y,t)=0`$ for all $`y>L(t)`$, this choice has no effect on our results. We now turn to a detailed investigation of our models. ## 2 Model A Firstly, we introduce a time dependent transition size, $`y_m(t)`$, between zero and the size of the largest particle in the system at time $`t`$, $`L(t)`$. The transition size produces two regions of fragment sizes with different rates of fragmentation above and below the transition size. At each time step there is a fixed probability, $`p_1`$, for the fragmentation of the largest particle in the system. A particle of size smaller than the transition size breaks with a probability $`p_2`$ and a rate proportional to the inverse of the fragment size, at each time step. Any other particle of size larger than the transition size can be fragmented, at each time step, either with a probability $`p_3`$ and a rate proportional to the inverse of the fragment size, or with a probability $`p_4`$ and a rate proportional to $`{\displaystyle \frac{\delta t}{y(t)}}\mathrm{ln}\left({\displaystyle \frac{y(t)}{\stackrel{~}{y}(t)}}\right)`$. Here $`y(t)`$ is the size of the fragment at time $`t`$ and $`\stackrel{~}{y}(t)`$ is a size for which the logarithmic rate is zero at time $`t`$. In the model we have $`p_1+p_2+p_3+p_4=1`$. We also assume that $`y_m(t)\stackrel{~}{y}(t)`$, such that for all fragments of sizes larger than the transition size, the logarithmic term in the rate of fragmentation is positive. In order to make the model tractable, we will later choose to have $`y_m(t)`$ proportional to $`L(t)`$. In the model the distribution function of daughter fragments is uniform. Therefore the rate of breaking of a particle into two smaller pieces is independent of the daughter fragment sizes. With probability $`p_2`$ every particle of size smaller than $`y_m(t)`$ is equally likely to be chosen for the fragmentation. With probabilities $`p_3`$ and $`p_4`$, every particle of size larger than the transition size is equally likely to be selected for the fragmentation with the correspondence rate. We assume that in time $`(t,t+\delta t)`$ the largest size changes from $`L(t)`$ to $`L(t)p_1\delta L`$ and the transition size changes from $`y_m(t)`$ to $`y_m(t)\delta y_m`$. Then, the fragmentation rate for this model is given by $`\delta tR(y,z)`$ $`=`$ $`p_1{\displaystyle \frac{\delta L}{L(t)}}\delta (y+zL(t))+p_2{\displaystyle \frac{\delta t}{y+z}}\theta (y_m(t)(y+z))`$ $`+\left[p_3{\displaystyle \frac{\delta t}{y+z}}+p_4{\displaystyle \frac{\delta t}{y+z}}\mathrm{ln}\left({\displaystyle \frac{y+z}{\stackrel{~}{y}(t)}}\right)\right]\theta ((y+z)y_m(t)).`$ The $`\delta `$-function in the first term on the right-hand side ensures that only the largest particle is fragmented and $`{\displaystyle \frac{\delta L}{L(t)}}`$ is the probability of placing the cut in the largest particle in a particular infinitesimal length $`\delta L`$. The second term on the right-hand side of (2) represents the rate of fragmentation for the particles of sizes smaller than $`y_m(t)`$ in time $`(t,t+\delta t)`$. The remaining terms on the right hand side of (2) represent the rate of fragmentation for the fragments of sizes between $`y_m(t)`$ and $`L(t)`$ in time $`(t,t+\delta t)`$. Inserting the rate (2) into the kinetic equation (1) for the fragments in both size regions smaller and larger than $`y_m(t)`$, yields $`n_1(y,t+\delta t)`$ $`=`$ $`n_1(y,t)p_2\delta tn_1(y,t)+\mathrm{\hspace{0.33em}2}p_1{\displaystyle \frac{\delta L}{L(t)}}n_2(L(t),t)`$ (3) $`+\mathrm{\hspace{0.17em}2}p_2\delta t{\displaystyle _y^{y_m(t)}}n_1(z,t){\displaystyle \frac{dz}{z}}+2p_3\delta t{\displaystyle _{y_m(t)}^{L(t)}}n_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.17em}2}p_4\delta t{\displaystyle _{y_m(t)}^{L(t)}}n_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{dz}{z}}`$ for $`y<y_m(t)`$ and $`n_2(y,t+\delta t)`$ $`=`$ $`n_2(y,t)p_3\delta tn_2(y,t)p_4\delta t\mathrm{ln}\left({\displaystyle \frac{y}{\stackrel{~}{y}(t)}}\right)n_2(y,t)`$ (4) $`+\mathrm{\hspace{0.25em}2}p_1{\displaystyle \frac{\delta L}{L(t)}}n_2(L(t),t)+\mathrm{\hspace{0.33em}2}p_3\delta t{\displaystyle _y^{L(t)}}n_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.25em}2}p_4\delta t{\displaystyle _y^{L(t)}}n_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{dz}{z}}`$ for $`y>y_m(t)`$. Here $`n_1(y,t)`$ and $`n_2(y,t)`$ are the densities of fragments of size $`y`$ at time $`t`$ which is smaller and larger than $`y_m(t)`$, respectively. The second term on the right hand side of (3) and the second and third terms on the right hand side of (4) represent the decrease in the number of particles of size $`y`$ from fragmentation into smaller particles, with probabilities $`p_2`$, $`p_3`$ and $`p_4`$, respectively. The next term on the right hand sides of (3,4) describes the gain in the number of particles of size $`y`$ from the fragmentation of the largest fragment of size $`L(t)`$. The remaining terms on the right hand sides of (3,4) are the increase in the number of particles of size $`y`$ from the fragmentation of all particles larger than $`y`$ and smaller than $`L(t)`$. Now we define densities of fragments of length $`y`$ at time $`t`$, which are normalised to a positive constant value A for two regions of sizes smaller and larger than $`y_m(t)`$, $$g_1(y,t)=A\frac{n_1(y,t)}{{\displaystyle _0^{y_m(t)}}n_1(z,t)𝑑z+{\displaystyle _{y_m(t)}^{L(t)}}n_2(z,t)𝑑z}$$ (5) for $`y<y_m(t)`$ and $$g_2(y,t)=A\frac{n_2(y,t)}{{\displaystyle _0^{y_m(t)}}n_1(z,t)𝑑z+{\displaystyle _{y_m(t)}^{L(t)}}n_2(z,t)𝑑z}$$ (6) for $`y>y_m(t)`$. In time $`tt+\delta t`$ as $`L(t)L(t)p_1\delta L`$ and $`y_m(t)y_m(t)\delta y_m`$ they yield $$g_1(y,t+\delta t)=A\frac{n_1(y,t+\delta t)}{{\displaystyle _0^{y_m(t)\delta y_m(t)}}n_1(z,t+\delta t)𝑑z+{\displaystyle _{y_m(t)\delta y_m(t)}^{L(t)p_1\delta L}}n_2(z,t+\delta t)𝑑z}$$ (7) for $`y<y_m(t)`$ and $$g_2(y,t+\delta t)=A\frac{n_2(y,t+\delta t)}{{\displaystyle _0^{y_m(t)\delta y_m(t)}}n_1(z,t+\delta t)𝑑z+{\displaystyle _{y_m(t)\delta y_m(t)}^{L(t)p_1\delta L}}n_2(z,t+\delta t)𝑑z}$$ (8) for $`y>y_m(t)`$. Inserting (3), (4) in (7), (8), respectively, gives us $`g_1(y,t+\delta t)=`$ (9) $`2p_1{\displaystyle \frac{\delta L}{L(t)}}g_2(L(t),t)+\mathrm{\hspace{0.17em}2}p_2\delta t{\displaystyle _y^{y_m(t)}}g_1(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.33em}2}p_3\delta t{\displaystyle _{y_m(t)}^{L(t)}}g_2(z,t){\displaystyle \frac{dz}{z}}+\mathrm{\hspace{0.17em}2}p_4\delta t{\displaystyle _{y_m(t)}^{L(t)}}g_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{dz}{z}}`$ $`+g_1(y,t)[1p_2\delta t{\displaystyle \frac{p_1}{A}}g_2(L(t),t)\delta L{\displaystyle \frac{p_2}{A}}\delta t{\displaystyle _0^{y_m(t)}}g_1(z,t)dz`$ $`{\displaystyle \frac{p_3}{A}}\delta t{\displaystyle _{y_m(t)}^{L(t)}}g_2(z,t)dz{\displaystyle \frac{p_4}{A}}\delta t{\displaystyle _{y_m(t)}^{L(t)}}g_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right)dz]`$ for $`y<y_m(t)`$ and $`g_2(y,t+\delta t)=`$ (10) $`2p_1{\displaystyle \frac{\delta L}{L(t)}}g_2(L(t),t)+\mathrm{\hspace{0.17em}2}p_3\delta t{\displaystyle _y^{L(t)}}g_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.33em}2}p_4\delta t{\displaystyle _y^{L(t)}}g_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{dz}{z}}+g_2(y,t)[1p_3\delta t`$ $`p_4\delta t\mathrm{ln}\left({\displaystyle \frac{y}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{p_1}{A}}g_2(L(t),t)\delta L{\displaystyle \frac{p_2}{A}}\delta t{\displaystyle _0^{y_m(t)}}g_1(z,t)𝑑z`$ $`{\displaystyle \frac{p_3}{A}}\delta t{\displaystyle _{y_m(t)}^{L(t)}}g_2(z,t)dz{\displaystyle \frac{p_4}{A}}\delta t{\displaystyle _{y_m(t)}^{L(t)}}g_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right)dz]`$ for $`y>y_m(t)`$. In order to solve (9) and (10) we introduce two functions $`F_1(x,t)`$ and $`F_2(x,t)`$ defined by $$F_1(x,t)=L(t)g_1(xL(t),t)$$ (11) for $`x<x_m`$ and $$F_2(x,t)=L(t)g_2(xL(t),t)$$ (12) for $`x>x_m`$. Here we have defined dimensionless and time independent variables $`x={\displaystyle \frac{y(t)}{L(t)}}`$, $`x_m={\displaystyle \frac{y_m(t)}{L(t)}}`$ and $`\stackrel{~}{x}={\displaystyle \frac{\stackrel{~}{y}(t)}{L(t)}}`$, where $`x`$ changes in the range $`[0,1]`$. $`x_m`$ and $`\stackrel{~}{x}`$ are fixed values and restricted to the same range. Using (9-12) we can obtain two partial differential equations for $`F_1(x,t)`$ and $`F_2(x,t)`$ as $`{\displaystyle \frac{F_1(x,t)}{t}}=`$ $`\nu x{\displaystyle \frac{F_1(x,t)}{x}}+2\nu F_2(1,t)+\mathrm{\hspace{0.17em}2}p_2{\displaystyle _x^{x_m}}F_1(z,t){\displaystyle \frac{dz}{z}}+2p_3{\displaystyle _{x_m}^1}F_2(x,t){\displaystyle \frac{dx}{x}}`$ $`+\mathrm{\hspace{0.33em}2}p_4{\displaystyle _{x_m}^1}F_2(x,t)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right){\displaystyle \frac{dx}{x}}F_1(x,t)[\nu +p_2+{\displaystyle \frac{\nu }{A}}F_2(1,t)`$ $`+{\displaystyle \frac{p_2}{A}}{\displaystyle _0^{x_m}}F_1(x,t)dx+{\displaystyle \frac{p_3}{A}}{\displaystyle _{x_m}^1}F_2(x,t)dx+{\displaystyle \frac{p_4}{A}}{\displaystyle _{x_m}^1}F_2(x,t)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)dx]`$ for $`x<x_m`$ and $`{\displaystyle \frac{F_2(x,t)}{t}}=`$ (14) $`\nu x{\displaystyle \frac{F_2(x,t)}{x}}+2\nu F_2(1,t)+\mathrm{\hspace{0.17em}2}p_3{\displaystyle _x^1}F_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.33em}2}p_4{\displaystyle _x^1}F_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{x}}}\right){\displaystyle \frac{dz}{z}}F_2(x,t)[\nu +p_3`$ $`+p_4\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)+{\displaystyle \frac{\nu }{A}}F_2(1,t)+{\displaystyle \frac{p_2}{A}}{\displaystyle _0^{x_m}}F_1(x,t)𝑑x`$ $`+{\displaystyle \frac{p_3}{A}}{\displaystyle _{x_m}^1}F_2(x,t)dx+{\displaystyle \frac{p_4}{A}}{\displaystyle _{x_m}^1}F_2(x,t)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)dx]`$ for $`x>x_m`$. Here we have set $`p_1{\displaystyle \frac{\delta L}{L(t)\delta t}}=\nu `$ to be a fixed value, and assumed that $`p_1>0`$. $`\nu `$ is equal to the rate of fragmentation for the largest fragment, and determines the relationship between the real time and the length of the largest fragment at time $`t`$, $`L(t)`$. In the limit $`t\mathrm{}`$, we assume that $`F_1(x,t)`$ and $`F_2(x,t)`$ evolve to time-independent quantities, $$F_1(x)=lim_t\mathrm{}F_1(x,t)$$ (15) for $`x<x_m`$ and $$F_2(x)=lim_t\mathrm{}F_2(x,t)$$ (16) for $`x>x_m`$. So that, as $`t\mathrm{}`$ then $`{\displaystyle \frac{F_1(x,t)}{t}}0`$ and $`{\displaystyle \frac{F_2(x,t)}{t}}0`$. To solve (2,14) in the long time limit using (15,16), we anticipate solutions as $$F_1(x)=Bx^\alpha $$ (17) for $`x<x_m`$ and $$F_2(x)=C\frac{\mathrm{exp}\left\{\beta \left[\mathrm{ln}\left(\frac{x}{\stackrel{~}{x}}\right)\right]^2\right\}}{x}$$ (18) for $`x>x_m`$. Substituting (17,18) in (2,14) in the long time limit, using (15,16) give us $$\alpha =\frac{p_2}{p_3},$$ (19) $$\beta =\frac{p_4}{2p_3}$$ (20) and $$\nu =p_3.$$ (21) B and C in (17),(18), respectively, are constant values and can be obtained using (2,14) in the long time limit and the continuity condition at $`x_m`$, $$F_1(x_m)=F_2(x_m).$$ (22) These give $$C=\frac{A}{\left({\displaystyle \frac{p_3}{p_3p_2}}\right)\mathrm{exp}\left\{\beta \left[\mathrm{ln}\left({\displaystyle \frac{x_m}{\stackrel{~}{x}}}\right)\right]^2\right\}+\sqrt{{\displaystyle \frac{2\pi p_3}{p_4}}}\left[P\left(\sqrt{{\displaystyle \frac{p_4}{p_3}}}\mathrm{ln}\left({\displaystyle \frac{1}{\stackrel{~}{x}}}\right)\right)P\left(\sqrt{{\displaystyle \frac{p_4}{p_3}}}\mathrm{ln}\left({\displaystyle \frac{x_m}{\stackrel{~}{x}}}\right)\right)\right]}$$ (23) and $$B=C\frac{\mathrm{exp}\left\{{\displaystyle \frac{p_4}{2p_3}}\left[\mathrm{ln}\left({\displaystyle \frac{x_m}{\stackrel{~}{x}}}\right)\right]^2\right\}}{x_m^{\left(1\frac{p_2}{p_3}\right)}},$$ (24) where the second and third terms in the denominator of (23) are a proportion of the normal probability distribution function defined as $$P(x)=\frac{1}{\sqrt{2\pi }}_{\mathrm{}}^xe^{t^2/2}𝑑t.$$ (25) We have also assumed that $$0\alpha =\frac{p_2}{p_3}<1.$$ (26) Solutions (17,18) using (19,20,23-26) satisfy the normalisation relation $$_0^{x_m}F_1(x)𝑑x+_{x_m}^1F_2(x)𝑑x=A.$$ (27) If $`A=1`$, the normalisation is equal to $`1`$. For a choice of $`A`$ as $$A=\sqrt{\frac{p_4}{2\pi p_3}}\frac{p_3}{p_3p_2}\mathrm{exp}\left\{\beta \left[\mathrm{ln}\left(\frac{x_m}{\stackrel{~}{x}}\right)\right]^2\right\}+P\left(\sqrt{\frac{p_4}{p_3}}\mathrm{ln}\left(\frac{1}{\stackrel{~}{x}}\right)\right)P\left(\sqrt{\frac{p_4}{p_3}}\mathrm{ln}\left(\frac{x_m}{\stackrel{~}{x}}\right)\right),$$ (28) the solutions (17,18) using (19,20,23,24) yield $$F_1(x)=\sqrt{\frac{p_4}{2\pi p_3}}\frac{\mathrm{exp}\left\{{\displaystyle \frac{p_4}{2p_3}}\left[\mathrm{ln}\left({\displaystyle \frac{x_m}{\stackrel{~}{x}}}\right)\right]^2\right\}}{x_m^{\left(1\frac{p_2}{p_3}\right)}}x^{\frac{p_2}{p_3}}$$ (29) for $`x<x_m`$ and $$F_2(x)=\sqrt{\frac{p_4}{2\pi p_3}}\frac{\mathrm{exp}\left\{{\displaystyle \frac{p_4}{2p_3}}\left[\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)\right]^2\right\}}{x}$$ (30) for $`x>x_m`$. We see that the stable distribution has a power law behaviour in the smaller region, $`x<x_m`$. For a log-normal distribution, the probability of finding a fragment of size between $`x`$ and $`x+dx`$ is given by $`n(x)dx`$, where $$n(x)=\frac{\mathrm{exp}\left\{\left[\mathrm{ln}\left(x/\stackrel{~}{x}\right)\right]^2/2\sigma ^2\right\}}{\left(2\pi \sigma ^2\right)^{1/2}x},$$ (31) where $`\stackrel{~}{x}`$ is the mean and $`\sigma `$ is the dispersion of the distribution. Comparing (30) with (31) shows that the stable distribution $`F_2(x)`$ in the larger region exhibits a log-normal behaviour with a mean and a dispersion of the distribution equal to $`\stackrel{~}{x}`$ and $`\sigma =\sqrt{{\displaystyle \frac{p_3}{p_4}}}`$, respectively. To have a positive value for $`A`$ in (28) and therefore physically meaningful distributions, we require $`p_3>p_2`$. This requirement is satisfied in (26). Now we consider some special cases of the model. 1. There is no fragmentation for the fragments smaller than $`y_m(t)`$, i.e. $`p_2=0`$ and $`p_1+p_3+p_4=1`$. The stable distribution then from (29) yields $$F_1(x)=\sqrt{\frac{p_4}{2\pi p_3}}\frac{\mathrm{exp}\left\{{\displaystyle \frac{p_4}{2p_3}}\left[\mathrm{ln}\left({\displaystyle \frac{x_m}{\stackrel{~}{x}}}\right)\right]^2\right\}}{x_m}$$ (32) for $`x<x_m`$ and is the same as (30) for $`x>x_m`$. We see that the distribution is a constant value in the smaller region and has a log-normal form in the larger region. 2. As $`y_m(t)=\stackrel{~}{y}(t)`$ and therefore $`x_m=\stackrel{~}{x}`$, the distributions (29,30) give $$F_1(x)=\sqrt{\frac{p_4}{2\pi p_3}}x_m^{\left(1\frac{p_2}{p_3}\right)}x^{\frac{p_2}{p_3}}$$ (33) for $`x<x_m`$ and $$F_2(x)=\sqrt{\frac{p_4}{2\pi p_3}}\frac{\mathrm{exp}\left\{{\displaystyle \frac{p_4}{2p_3}}\left[\mathrm{ln}\left({\displaystyle \frac{x}{x_m}}\right)\right]^2\right\}}{x}$$ (34) for $`x>x_m`$. Again the distribution reveals a power law form in the smaller region and has a log-normal behaviour in the larger region. 3. As $`p_40`$ and therefore $`\sigma ^2={\displaystyle \frac{p_3}{p_4}}\mathrm{}`$, the distribution (30) in the larger region yield $`F_2(x)x^{\mathrm{\hspace{0.17em}1}}`$. On the other hand as $`p_40`$ the rate of fragmentation (2) for the fragments of sizes larger than $`y_m(t)`$ is only proportional to the inverse of the fragment size. Consequently, at each time step either the largest fragment breaks with probability $`p_1`$, or a fragment smaller than $`y_m(t)`$ is chosen for fragmentation with probability $`p_2`$, or a fragment larger than $`y_m(t)`$ is selected with probability $`p_3`$, where $`p_1+p_2+p_3=1`$. The stable distributions then exhibit a composite power law behaviour with power laws of exponents equal to $`{\displaystyle \frac{p_2}{\nu }}`$ and $`{\displaystyle \frac{p_3}{\nu }}`$ in the smaller and larger region, respectively. In our case since from (21) we have $`\nu =p_3`$, the exponents reduce to $`{\displaystyle \frac{p_2}{p_3}}`$ and $`1`$ in the two regions. The power law distribution in the smaller region is similar to that from (17,19) and in the larger region has an exponent of $`1`$ as expected. The latter is consistent with one of interesting properties of the log-normal distribution where as the variance $`\sigma ^2`$ is pretty large, then the distribution can be approximated by a power law form with an exponent equal to $`1`$, . ## 3 Model B In this model again a time dependent transition size, $`y_m(t)`$, produces two regions of fragment sizes with different rates of fragmentation above and below the transition size. At each time step either the largest fragment in the system breaks with a fixed probability, $`p_1`$, or a fragment smaller than the transition size is chosen for the fragmentation with probability $`p_2`$ and a rate proportional to the inverse of fragment size, or another fragment larger than $`y_m(t)`$ is selected with probability $`(1p_1p_2)`$ and a rate proportional to $`{\displaystyle \frac{\delta t}{y(t)}}\left[1+\lambda \mathrm{ln}\left({\displaystyle \frac{y(t)}{\stackrel{~}{y}(t)}}\right)\right]`$. Here $`y(t)`$ is the size of fragment at time $`t`$, $`\lambda `$ is a positive and constant value and $`\stackrel{~}{y}(t)`$ is a size for which the logarithmic term is zero at time $`t`$. In this model we assume that $`y_m(t)<\stackrel{~}{y}(t)`$. For a fragment of size $`y>\stackrel{~}{y}(t)`$ the logarithmic term in the rate of fragmentation is always positive, but a fragment of size $`y_m(t)<y<\stackrel{~}{y}(t)`$ gives a negative value for the logarithmic term. In order to have a positive rate of fragmentation for all particles larger than $`y_m(t)`$, we require $`\left[1+\lambda \mathrm{ln}\left({\displaystyle \frac{y(t)}{\stackrel{~}{y}(t)}}\right)\right]>0`$ which gives $`y(t)>\stackrel{~}{y}(t)e^{\mathrm{\hspace{0.17em}1}/\lambda }`$. If we choose $`y_m(t)`$ to be $$y_m(t)=\stackrel{~}{y}(t)e^{\mathrm{\hspace{0.17em}1}/\lambda },$$ (35) then the rate of fragmentation for all fragments of sizes larger than $`y_m(t)`$ is positive. Consequently, the rate of fragmentation for this model is given by $`\delta tR(y,z)`$ $`=`$ $`p_1{\displaystyle \frac{\delta L}{L(t)}}\delta (y+zL(t))+p_2{\displaystyle \frac{\delta t}{y+z}}\theta (y_m(t)(y+z))`$ $`+(1p_1p_2){\displaystyle \frac{\delta t}{y+z}}\left[1+\lambda \mathrm{ln}\left({\displaystyle \frac{y+z}{\stackrel{~}{y}(t)}}\right)\right]\theta ((y+z)y_m(t)),`$ where $`y_m(t)`$ is defined as (35). The first two terms on the right hand side of (3) are the same as those in (2) and have the same definitions as before. The last term represents the rate of fragmentation for the fragments of sizes larger than $`y_m(t)`$ in time $`(t,t+\delta t)`$. Using (1,3) for the both size regions give the kinetic equations $`n_1(y,t+\delta t)=`$ (37) $`n_1(y,t)p_2\delta tn_1(y,t)+\mathrm{\hspace{0.33em}2}p_1{\displaystyle \frac{\delta L}{L(t)}}n_2(L(t),t)`$ $`+\mathrm{\hspace{0.25em}2}p_2\delta t{\displaystyle _y^{y_m(t)}}n_1(z,t){\displaystyle \frac{dz}{z}}+2(1p_1p_2)\delta t{\displaystyle _{y_m(t)}^{L(t)}}n_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.25em}2}\lambda (1p_1p_2)\delta t{\displaystyle _{y_m(t)}^{L(t)}}n_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{dz}{z}}`$ for $`y<y_m(t)`$ and $`n_2(y,t+\delta t)=`$ (38) $`n_2(y,t)(1p_1p_2)\delta tn_2(y,t)\lambda (1p_1p_2)\delta t\mathrm{ln}\left({\displaystyle \frac{y}{\stackrel{~}{y}(t)}}\right)n_2(y,t)`$ $`+\mathrm{\hspace{0.25em}2}p_1{\displaystyle \frac{\delta L}{L(t)}}n_2(L(t),t)+\mathrm{\hspace{0.25em}2}(1p_1p_2)\delta t{\displaystyle _y^{L(t)}}n_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.25em}2}\lambda (1p_1p_2)\delta t{\displaystyle _y^{L(t)}}n_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{y}(t)}}\right){\displaystyle \frac{dz}{z}}`$ for $`y>y_m(t)`$. Following the same procedure as that used in the model A, we obtain $`{\displaystyle \frac{F_1(x,t)}{t}}=`$ $`\nu x{\displaystyle \frac{F_1(x,t)}{x}}+2\nu F_2(1,t)+\mathrm{\hspace{0.17em}2}p_2{\displaystyle _x^{x_m}}F_1(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.33em}2}(1p_1p_2){\displaystyle _{x_m}^1}F_2(x,t){\displaystyle \frac{dx}{x}}+\mathrm{\hspace{0.17em}2}\lambda (1p_1p_2){\displaystyle _{x_m}^1}F_2(x,t)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right){\displaystyle \frac{dx}{x}}`$ $`F_1(x,t)[\nu +p_2+{\displaystyle \frac{\nu }{A}}F_2(1,t)+{\displaystyle \frac{p_2}{A}}{\displaystyle _0^{x_m}}F_1(x,t)dx`$ $`+{\displaystyle \frac{(1p_1p_2)}{A}}{\displaystyle _{x_m}^1}F_2(x,t)dx+{\displaystyle \frac{\lambda (1p_1p_2)}{A}}{\displaystyle _{x_m}^1}F_2(x,t)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)dx]`$ for $`x<x_m`$ and $`{\displaystyle \frac{F_2(x,t)}{t}}=`$ $`\nu x{\displaystyle \frac{F_2(x,t)}{x}}+2\nu F_2(1,t)+\mathrm{\hspace{0.17em}2}(1p_1p_2){\displaystyle _x^1}F_2(z,t){\displaystyle \frac{dz}{z}}`$ $`+\mathrm{\hspace{0.25em}2}\lambda (1p_1p_2){\displaystyle _x^1}F_2(z,t)\mathrm{ln}\left({\displaystyle \frac{z}{\stackrel{~}{x}}}\right){\displaystyle \frac{dz}{z}}F_2(x,t)[\nu +(1p_1p_2)`$ $`+\lambda (1p_1p_2)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)+{\displaystyle \frac{\nu }{A}}F_2(1,t)+{\displaystyle \frac{p_2}{A}}{\displaystyle _0^{x_m}}F_1(x,t)𝑑x`$ $`+{\displaystyle \frac{(1p_1p_2)}{A}}{\displaystyle _{x_m}^1}F_2(x,t)dx+{\displaystyle \frac{\lambda (1p_1p_2)}{A}}{\displaystyle _{x_m}^1}F_2(x,t)\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)dx]`$ for $`x>x_m`$. Here $`x`$, $`x_m`$, $`\stackrel{~}{x}`$ and $`\nu `$ have the same definitions as before and from (35) we get $$x_m=\stackrel{~}{x}e^{\mathrm{\hspace{0.17em}1}/\lambda }.$$ (41) In the long time limit as $`t\mathrm{}`$, we assume that $`F_1(x,t)`$ and $`F_2(x,t)`$ evolve to the same time-independent variables as (15) and (16), respectively. In order to solve (3,3) in this limit we anticipate solutions as $$F_1(x)=Gx^\gamma $$ (42) for $`x<x_m`$ and $$F_2(x)=H\frac{\mathrm{exp}\left\{\eta \left[\mathrm{ln}\left(\frac{x}{\stackrel{~}{x}}\right)\right]^2\right\}}{x}$$ (43) for $`x>x_m`$. Substituting (42,43) in (3,3) in the long time limit, gives us $$\gamma =\frac{p_2}{1p_1p_2},$$ (44) $$\eta =\frac{\lambda }{2}$$ (45) and $$\nu =\mathrm{\hspace{0.33em}1}p_1p_2.$$ (46) G and H in (42,43) are constant values and can be obtained using (3,3) in the long time limit and the continuity condition at $`x_m`$, (22). These give $$H=\frac{A}{{\displaystyle \frac{1p_1p_2}{1p_12p_2}}\mathrm{exp}\left({\displaystyle \frac{1}{2\lambda }}\right)+\sqrt{{\displaystyle \frac{2\pi }{\lambda }}}\left[P\left(\sqrt{\lambda }\mathrm{ln}\left({\displaystyle \frac{1}{\stackrel{~}{x}}}\right)\right)P\left({\displaystyle \frac{1}{\sqrt{\lambda }}}\right)\right]}$$ (47) and $$G=H\frac{\mathrm{exp}\left({\displaystyle \frac{1}{2\lambda }}\right)}{\left(\stackrel{~}{x}e^{\mathrm{\hspace{0.17em}1}/\lambda }\right)^{\frac{1p_12p_2}{1p_1p_2}}},$$ (48) where we have assumed that $$0\gamma =\frac{p_2}{1p_1p_2}<1.$$ (49) The second and third terms in the denominator of (47) are a proportion of the normal probability distribution function with the same definition as (25). Solutions (42,43) using (44-49) satisfy the normalisation relation (27). If $`A=1`$, the normalisation is equal to $`1`$. For a choice of $`A`$ as $$A=\sqrt{\frac{\lambda }{2\pi }}\frac{1p_1p_2}{1p_12p_2}\mathrm{exp}\left(\frac{1}{2\lambda }\right)+P\left(\sqrt{\lambda }\mathrm{ln}\left(\frac{1}{\stackrel{~}{x}}\right)\right)P\left(\frac{1}{\sqrt{\lambda }}\right),$$ (50) the solutions (42,43) using (44,45,47,48) yield $$F_1(x)=\sqrt{\frac{\lambda }{2\pi }}\frac{\mathrm{exp}\left({\displaystyle \frac{1}{2\lambda }}\right)}{\left(\stackrel{~}{x}e^{\mathrm{\hspace{0.17em}1}/\lambda }\right)^{\frac{1p_12p_2}{1p_1p_2}}}x^{\frac{p_2}{1p_1p_2}}$$ (51) for $`x<\stackrel{~}{x}e^{\mathrm{\hspace{0.17em}1}/\lambda }`$ and $$F_2(x)=\sqrt{\frac{\lambda }{2\pi }}\frac{\mathrm{exp}\left\{{\displaystyle \frac{\lambda }{2}}\left[\mathrm{ln}\left({\displaystyle \frac{x}{\stackrel{~}{x}}}\right)\right]^2\right\}}{x}$$ (52) for $`x>\stackrel{~}{x}e^{\mathrm{\hspace{0.17em}1}/\lambda }`$. We see that the stable distribution has a power law behaviour in the smaller region. Comparing (52) with (31) exhibits that the stable distribution $`F_2(x)`$ in the larger region has a log-normal form with a mean of $`\stackrel{~}{x}`$ and a dispersion of the distribution equal to $$\sigma =\frac{1}{\sqrt{\lambda }}.$$ (53) It is shown that the log-normal function (52) has a maximum at $`x=\stackrel{~}{x}e^{\mathrm{\hspace{0.17em}1}/\lambda }`$ which we have assumed to be equal to $`x_m`$. This shows that in this model the dimensionless transition size $`x_m`$ is chosen to be equal to the value for which the log-normal distribution in the larger region is maximum. To have a positive value for $`A`$ in (50) and therefore physically meaningful distributions, it is required that $`(1p_1p_2)>p_2`$. This inequality is satisfied from (49). Now we consider some special cases of the model B. 1. When there is no fragmentation for the fragments smaller than $`y_m(t)`$, i.e. $`p_2=0`$, then the stable distribution from (51) yields $$F_1(x)=\sqrt{\frac{\lambda }{2\pi }}\frac{1}{\stackrel{~}{x}}e^{1/2\lambda }$$ (54) for $`x<x_m`$ and is the same as (52) for $`x>x_m`$. The distribution is a constant value in the smaller region and has a log-normal behaviour in the larger region. 2. As $`\lambda 0`$, and therefore from (53) $`\sigma ^2\mathrm{}`$, the situation is then similar to the case $`3`$ of model A and the log-normal distribution (52) in the larger region can be approximated by a power law form with an exponent equal to $`1`$, $`F_2(x)x^{\mathrm{\hspace{0.17em}1}}`$. On the other hand at this limit the rate of fragmentation (3) gives two different probabilities in the two different size regions with a rate proportional to the inverse of the fragment size in each region. The stable distribution then reveal a composite power law with power laws of exponents of $`{\displaystyle \frac{p_2}{\nu }}`$ and $`{\displaystyle \frac{1p_1p_2}{\nu }}`$ in the smaller and larger size regions, respectively. Since in our case from (46) we have $`\nu =(1p_1p_2)`$, the exponents yield $`{\displaystyle \frac{p_2}{1p_1p_2}}`$ and $`1`$, respectively. The result in the smaller size region is equivalent to that from (42,44). In the larger size region, as expected, we get a power law with an exponent of $`1`$ which is consistent with the property of a log-normal distribution with a very large variance. We see that all the results for the model B can be recovered from the model A by substituting $`p_3`$ by $`(1p_1p_2)`$, $`p_4`$ by $`\lambda (1p_1p_2)`$ and using equation (35). ## 4 Conclusions In this paper we have studied two models of binary fragmentation. In the models we introduced a time dependent transition size, $`y_m(t)`$, which produced two regions of fragment sizes with different rates of fragmentation above and below the transition size. The rate of fragmentation at each time step is fixed for the largest fragments and is proportional to the inverse of fragment size for the fragments smaller than the transition size. We considered a rate of fragmentation combined of two terms, one proportional to the inverse of the fragment size and the other proportional to a logarithmic function of the fragment size, for the fragments larger than the transition size at each time step. For a size equal to $`\stackrel{~}{y}(t)`$ at time $`t`$, the logarithmic function is zero. In model A we assumed that the transition size, $`y_m(t)`$ is any arbitrary value for which $`y_m(t)\stackrel{~}{y}(t)`$, whereas in the model B we chose a value as (35) for the transition size. The models were then solved exactly in the long time limit to reveal stable time-invariant solutions for the distributions. The results of these distributions for both models A and B exhibited a power law form with an exponent between $`1`$ and zero, in the smaller region. For a special choice of normalisation, the distributions exhibited log-normal behaviours in the larger regions. In this region they revealed one of interesting properties of a log-normal distribution for which as the variance of the log-normal is pretty large, then the distribution is approximated by a power law form with an exponent equal to $`1`$. Special cases of the models with no fragmentation for the smaller fragments were also examined. The stable distributions then exhibited a constant value and a log-normal form in the smaller and larger regions, respectively. Now we investigate some experimental work on the fragmentation involving log-normal distributions and the transition of the distributions from a log-normal form to a power law one and vice versa. In the experimental results of fragment size and mass distributions for the long thin glass rods with fixed lengths of $`1500mm`$ and diameters of $`2mm`$ were reported. The rods were dropped horizontally onto a flat hard floor from different heights. At lower falling heights (about $`1m`$ drop), the distributions exhibited a log-normal form for larger fragments and a power law form for smaller ones. The crossover was seen to be at length scales around the rod diameter. This is due to the fact that fragments smaller than the rod diameter, undergo a three-dimensional fracture, whereas the larger fragments are produced by one-dimensional breaking. For a falling height of $`1.20m`$ the cumulative number of fragments on their size were plotted. The data was fitted to a curve calculated by assuming that the size distribution could be described by a log-normal distribution. The fitting gave values of $`\stackrel{~}{l}=30mm`$ and $`\sigma =0.55`$ for the mean and the dispersion of the log-normal distribution. The fitting is excellent for the fragments larger than $`l_c=7mm`$ which is a length scale around the rod diameter. The size distribution of fragments smaller than $`l_c`$ seemed to have a power law form. As the falling height was increased, the cumulative number variation for larger fragments started to show a power law dependence on their size and mass. In experiments in rupture of mercury droplets were performed. In the experiment at a height of $`h`$, the mercury droplets of about $`2.0mm`$ radius were fallen directly on a glass Petri dish. The results of the cumulative number of the drops versus the drop diameter were plotted. For small falling heights a log-normal behaviour was exhibited. The distribution showed a clear transition from a log-normal form to a scaled one as the falling height $`h`$ was increased. Our models in the long time limit can be applied to the experimental systems when the fragmentation process is over and the system reaches a stable and time-independent phase. In the models we assumed that the fragments have only one dimension, whereas in real physical systems the fragments have three dimensions. We can therefore only apply our models to the systems for which the fractures occur in one dimension. This assumption seems physically reasonable for a long thin glass rod for which most of the fractures occur in the length of the rod and the cross-section remains almost fixed. We can also consider this assumption for the mercury droplets for which at each time step only the radius of the droplet changes and the shape of the fragments remains always spherical. Now we discuss the similarities between the results from our models and those from the experiments. The experimental results for the size distribution of long thin glass rods at low falling heights exhibited a power law and log-normal behaviour in the smaller and larger size regions, respectively . We got the same behaviour for the distributions in both our models. The data from the experiments for the falling height of $`1.20m`$ revealed that $`l_c<\stackrel{~}{l}`$, where $`l_c`$ is the transition size and $`\stackrel{~}{l}`$ is the mean of the log-normal distribution. In the model B of our work also we have $`x_m<\stackrel{~}{x}`$. We therefore expect model B of our work to describe the experimental results. In model B, the dispersion of the log-normal distribution can be obtained from (53) and the mean and the transition values, $`\stackrel{~}{x}`$ and $`x_m`$, respectively, are any arbitrary values which satisfy the relation (41). Then, by appropriate choices of $`\stackrel{~}{x}`$ and $`\lambda `$ in the model B of our work we can match the mean and the dispersion of the log-normal distribution in the experimental results. Because of (41,53) we should expect a relation as $`l_c=\stackrel{~}{l}e^{\sigma ^2}`$. The values of $`l_c`$ and $`\stackrel{~}{l}`$ from experimental result do not satisfy this relation. As a result, although model B of our work does not match the exact relation between $`l_c`$ and $`\stackrel{~}{l}`$ from experimental result, but it can give an explanation for the transition of distribution from a power law to a log-normal form on the shock fragmentation of long thin glass rods . Also in the smaller region of the model B, the appropriate choices of $`p_1`$ and $`p_2`$ can match the exponent of power law equal to $`{\displaystyle \frac{p_2}{(1p_1p_2)}}`$ from (42,44), with the corresponding exponent of power law distribution from the experimental results. This exponent needs to satisfy the condition (49). For high falling heights a transition of the distribution from log-normal to scaling law was observed in the experiments for the larger rod fragments and for the mercury droplets . At high heights therefore the distributions from the experiments exhibit a composite and a single power law form for the rod fragments and mercury droplets, respectively. The models in this paper suggest a power law and a log-normal behaviour for the distribution of the smaller and larger fragments, respectively. To explain composite and single power law distributions, we can however use the models in . To conclude we introduced statistical models with a fixed probability for the breaking the largest particle in the system. All other particles of sizes smaller than the largest size, were divided into two different regions with different rates of fragmentation in each region. In the models therefore the largest fragments are broken with very large probability. This fact together with the form of the fragmentation rate as a function of the fragment size in each region, give a distribution of power law or log-normal form in that region. Our models therefore give a theoretical explanation for the transition of distribution from the scaling to log-normal behaviour, which had been seen before in the experiments. Acknowledgement ZT is grateful to the Soudavar Scholarship Fund for the award of a scholarship.
warning/0003/hep-lat0003019.html
ar5iv
text
# Measuring the aspect ratio renormalization of anisotropic-lattice gluons ## 1 Introduction In principle, an improved action makes it possible to achieve lattice volumes large enough to overcome finite size effects at a computational effort low enough to obtain measurements with good statistical errors. However typical masses for heavy states may be similar to or larger than the inverse lattice spacing on the associated coarse lattice. Propagators for such states then decay too fast in lattice units for accurate measurement. The problem can be overcome by tuning couplings associated with the “time” direction so that the temporal lattice spacing $`a_t`$ is much smaller than the spatial lattice spacing $`a_s`$ . Refined measurements are then possible while retaining the computational advantages of the improved action and coarse spatial lattice. Such anisotropic actions have already been successfully applied to the glueball spectrum , the spectrum of excitations of the inter-quark potential , heavy hybrids , and the fine structure of the quarkonium spectrum . They are also expected to be a powerful tool in extracting excited-state signals, obtaining high-momentum form factors, pushing thermodynamic calculations to higher temperatures, and in the calculation of transport coefficients. In this paper, we consider one formulation of the anisotropic action for the pure Yang-Mills sector : $`S_n=`$ $`\beta {\displaystyle \underset{x,s>s^{}}{}}\chi _0^1\left\{{\displaystyle \frac{5}{3}}{\displaystyle \frac{P_{s,s^{}}}{u_s^4}}{\displaystyle \frac{1}{12}}{\displaystyle \frac{R_{ss,s^{}}}{u_s^6}}{\displaystyle \frac{1}{12}}{\displaystyle \frac{R_{s^{}s^{},s}}{u_s^6}}\right\}\beta {\displaystyle \underset{x,s}{}}\chi _0\left\{{\displaystyle \frac{4}{3}}{\displaystyle \frac{P_{s,t}}{u_s^2u_t^2}}{\displaystyle \frac{1}{12}}{\displaystyle \frac{R_{ss,t}}{u_s^4u_t^2}}\right\}.`$ Here $`s,s^{}`$ run over spatial directions, $`P_{s,s^{}}`$ is a $`1\times 1`$ plaquette and $`R_{s^{}s^{},s}`$ is a $`2\times 1`$ rectangle. The coefficients of these terms are chosen so the action has no $`O(a_s^2)`$ discretization errors in tree-level pertubation theory. $`\chi _0`$ is the anisotropy parameter and is equal to the aspect ratio of the spatial and temporal lattice spacings, $`a_s`$ and $`a_t`$ at tree-level. At higher orders in the perturbative expansion, this aspect ratio recieves quantum corrections, so a renormalized anisotropy determined from a physical process, $`\chi _R`$, differs from $`\chi _0`$ at $`O(\alpha _s)`$. Tadpole improvement (TI) of the perturbative expansion is achieved by tuning the input spatial and temporal link parameters $`u_s`$ and $`u_t`$ for self-consistency at each choice of $`\beta `$ and $`\chi _0`$. Here we report on the accurate determination of the tadpole parameters in the plaquette and Landau mean-link formulations. (Non-tadpole-improved actions have been studied in ref. .) We compare methods for measuring the renormalized anisotropy and discuss the importance of lattice artifacts. In order that the anisotropic formulation can be used with confidence, physically distinct methods for determining the renormalized anisotropy, $`\chi _R=a_s/a_t`$, should agree. Discrepancies in the results for $`\chi _R`$ give a measure of lattice artifacts and an indication of the importance of further improvements. Our results have been used in applications of anisotropic lattices to heavy quark hybrid states . In section 2 we report on the determination of the tadpole parameters; in section 3 the renormalized anisotropy measurements are presented, in section 5 we discuss the results and present conclusions. ## 2 Tadpole Parameters In both the plaquette and Landau mean-link formulations, the tadpole parameters are determined self-consistently and are defined by $$\begin{array}{ccccccccc}u_s\hfill & =& \left(P_{s,s^{}}\right)^{1/4}\hfill & ,\hfill & u_t\hfill & =& 1\hfill & & \text{plaquette},\hfill \\ & & & & & & & & \\ u_s\hfill & =& U_s_{Landau}\hfill & ,\hfill & u_t\hfill & =& U_t_{Landau}\hfill & & \text{Landau},\hfill \end{array}$$ (2.1) where the Landau gauge is defined by the field configuration which maximizes $$F(\{U\})=\underset{x\mu }{}\frac{1}{u_\mu a_\mu ^2}\text{ReTr}\left\{U_\mu (x)\frac{1}{16u_\mu }U_\mu (x)U_\mu (x+\widehat{\mu })\right\}.$$ (2.2) with respect to gauge transformations. We denote the gauge coupling in the two schemes as $`\beta _L`$ and $`\beta _P`$ respectively. ### 2.1 Plaquette Tadpoles In the plaquette scheme of Eqn. (2.1), $`u_t=1`$ and $`u_s`$ is determined self-consistently. The expectation value of the spatial plaquette is computed for a range of input parameters $`u_s`$ close to and spanning the self-consistent value, $`u_s^{}`$. A linear interpolation is sufficient to give an accurate value of $`u_s^{}`$. This value is then checked in a further Monte-Carlo simulation. The values of $`u_s^{}`$ for a set of couplings $`\beta _P`$ and tree-level anisotropies $`\chi _0`$ are given in Table 1. For the lattice sizes used in simulation, the plaquette expectation value is found to be independent of the volume at the four-significant-figure level. An extrapolation to infinite volume is unnecessary. ### 2.2 Landau Gauge Fixing The maximization of $`F(U_\mu ^g)`$ in Eqn. (2.2), where $`U_\mu ^g=g(x)U_\mu (x)g^{}(x+\mu )`$ with respect to a gauge transformation $`\{g(x)\}`$, was carried out using the conjugate-gradient method modified to deal with the group structure of the link elements. At each stage of the iteration the the appropriate conjugate-gradient vector $`\{𝒗(x)\}`$, is computed as a covariant derivative . $$𝒗(x)=\frac{}{\eta (x)}F(U_\mu ^g)|_{\eta =0},$$ (2.2.1) where $`g(x)=e^{\eta (x).T}`$ and $`T`$ are the generators of $`SU(3)`$ . The vector $`\{𝒗(x)\}`$ has elements lying in the Lie algebra of $`SU(3)`$ . For each vector a series of group elements is constructed in the associated one-parameter subgroup, each element being a given power of the preceding one: $$g_1=\mathrm{exp}(ϵ𝒗),g_n=(g_{n1})^p,1<nN.$$ (2.2.2) The local maximum in the direction $`𝒗`$ is found by evaluating $`F(U_\mu ^{g_p})`$ in descending order from $`p=N`$ to $`p=1`$ . The value of $`ϵ`$, at any stage, is reduced until either the sequence exhibits a maximum or remains constant within a preset tolerance. For large $`\chi _0`$ the terms including $`U_t`$ dominate the expression for $`F`$ and near to the maximum relatively small changes in $`F`$ correspond to quite large changes in $`u_s`$. Consequently, to be sure that $`u_s`$ as well as $`u_t`$ is accurately calculated, the maximum must be found to a sufficiently high accuracy. The criterion chosen was that $`\delta F/F2\times 10^6`$ where $`\delta F`$ was the accumulated change in $`F`$ for three consecutive iterations of the conjugate-gradient algorithm for which the individual changes in $`F`$ were non-zero. This criterion accounts for the observation that from time to time close to the maximum the change in $`F`$ was zero. We chose $`ϵ=2\times 10^5`$, $`p=4`$, and $`N=6`$. ### 2.3 Landau Gauge Fixed Tadpoles The self-consistent value of $`𝒖=(u_s,u_t)`$ can be determined by a generalized Newton-Rapheson method or by linear interpolation. The latter method was used for the mean-link tadpoles and was implemented by choosing four input tadpoles $`𝒖_i,i=1,\mathrm{},4`$ and measuring the four corresponding output tadpoles $`𝒖_i^{},i=1,4`$. A linear map is assumed for the incremental vectors $`𝐫_i=(𝒖_i𝒖_1),𝐫_i^{}=(𝒖_i^{}𝒖_1^{}),i=2,3,4`$: $$𝐫_i=𝑴𝐫_i^{},i=2,3,4,$$ (2.3.1) where $`𝑴`$ is a $`2\times 2`$ matrix that is determined from the the measured images of $`𝐫_2,𝐫_3`$ and checked for consistency against the measured image of $`𝐫_3`$ . The self-consistent tadpole $`𝒖^{}`$ is then predicted to be $$𝒖^{}=𝒖_1+(1𝑴)^1(𝐫_1^{}𝐫_1).$$ (2.3.2) The self-consistency of $`𝒖^{}`$ is then checked computationally. This method was found to be very reliable and only subject to minor adjustments to account for statistical errors. Moreover, a meaningful statistical error can be assigned to $`𝒖^{}`$ from the image under $`M^1`$ of the statistical error box deduced for $`𝒖^{}`$ from the measurements values of the tadpole parameters. Let the image of $`𝒖^{}`$ under the map $`M`$ be $`𝒖^{}`$ with statistical error $`\delta 𝒖^{}`$. The value $`𝒖^{}`$ is acceptable if in terms of components $`𝒖^{}\delta 𝒖^{}<𝒖^{}<𝒖^{}+\delta 𝒖^{}`$, and the error quoted on $`𝒖^{}`$ is $`\delta 𝒖^{}=M^1\delta 𝒖^{}`$. A typical example is for $`\beta _L=1.8,\chi _0=4`$ on a $`6^3\times 24`$ lattice where $$𝑴^1=\left(\begin{array}{cc}\hfill 0.341& \hfill 0.227\\ \hfill 0.033& \hfill 0.957\end{array}\right),\delta 𝒖^{}=\left(\begin{array}{c}\hfill 3\times 10^4\\ \hfill 3\times 10^5\end{array}\right)\delta 𝒖^{}=\left(\begin{array}{c}\hfill 1\times 10^4\\ \hfill 2\times 10^5\end{array}\right).$$ (2.3.3) In practice it is found that the error on $`u_s^{}`$ is typically reduced by a factor of 3 compared with the statistical error from the simulation verifying the self-consistency. The simulation for the mean-link tadpoles was done on the Hitachi SR2201 computers at the Cambridge High Performance Computing Facility and the Tokyo University Computer Centre. The lattices used were $`L^3\times T`$ with $`T=\chi _0L`$ and $`L=6,8`$ in all cases except one. The high accuracy demanded by the maximization process was very time consuming. Also, many configurations were required to achieve the desired statistical accuracy. Consequently, only one example with $`L=10`$ was done as a check on the finite-size scaling ansatz, $`𝒖^{}(L)=𝒖^{}(\mathrm{})+A/L^2`$ . Typically, for $`10^3\times 40`$ about 400 conjugate-gradient iterations were needed taking about 60 seconds per iteration. In the cases $`L=6,8,10`$ for $`\beta _L=1.8\chi _0=4`$, the fit to the finite-size scaling ansatz was very good as can be seen in Figs. 1 and 2. For other cases the linear extrapolation was assumed to hold. The values measured for $`𝒖^{}(L)`$ and the extrapolation to $`L=\mathrm{}`$ are given in table 2. ## 3 Anisotropy In this section we report on two methods for determining $`\chi _R`$. The first uses the dispersion relation for the torelon and the second uses the comparison of the potential measured in the fine and coarse direction using Wilson loops. ### 3.1 The Torelon The lattice considered is $`S^2\times L\times T`$ with typical values $`S=8`$ in the $`x`$ and $`y`$ directions, $`L=3,4,5`$ in the $`z`$ direction, and $`T=50`$ in the fine or $`t`$ direction. The lattice has periodic boundary conditions and the torelon is created by a Polyakov line that loops around the lattice in the $`z`$ direction. The Polyakov line is associated with a particular point in the $`(x,y)`$-plane and is constructed from links which are covariantly smeared in a manner similar to APE smearing: $`T(x,y,t)`$ $`=`$ $`Tr{\displaystyle \underset{z=0}{\overset{z=L1}{}}}W_z(x,y,z,t),`$ $`W_z(x,y,z,t)`$ $`=`$ $`(1+{\displaystyle \frac{l^2D^2}{4m}})^mU_z(x,y,z,t),`$ (3.1.1) where $`D`$ is the appropriate covariant derivative. Typically, $`l=1.5`$, $`m=10`$. The state with momentum $`𝒑=(p_x,p_y)=(n_x,n_y)(2\pi /Sa_s)`$ is $$T(𝒑,t)=\underset{x,y}{}T(x,y,t)e^{i(p_xx+p_yy)}.$$ (3.1.2) The torelon propagator $$G_T(𝒑,t)=\frac{1}{T}\underset{t^{}=0}{\overset{t^{}=T1}{}}T(𝒑,t^{})T^{}(𝒑,t+t^{}),$$ (3.1.3) is measured for various choices of momentum $`𝒑`$ and a correlated simultaneous fit using SVD decomposition is made to the relativistic dispersion formula: $`G_T(𝒑,t)`$ $`=`$ $`c(𝒑)e^{E(𝐩)a_t\overline{t}},`$ $`E(𝒑)a_t`$ $`=`$ $`{\displaystyle \frac{a_s\sqrt{𝒑^2+M_T^2}}{\chi _R}},`$ (3.1.4) where $`M_T`$ is the torelon mass and $`\overline{t}=t/a_t`$. The momenta used were $`𝒏^2=n_x^2+n_y^2=0,1,2,4,5`$ with $`S=8`$. An example of the fit obtained is shown in Fig. 3 for $`\beta _L=1.8,\chi _0=6`$. The fit is over the range $`2t12`$ and has $`\chi ^2/N_{\mathrm{df}}=0.96`$ for 43 d.o.f. The fit is good and gives $`(M_Ta_s)^2=1.85(4),\chi _R=3.61(2)`$. In Fig. 4 the dispersion curves $`E(𝒑^2)`$ versus $`𝒏^2`$ are plotted for $`\beta _L=1.8,\chi _0=4,L=3,4,5`$. From both Figs. 3 and 4 it is clear that rotational invariance is established in the coarse $`xy`$ directions. The good simultaneous fit to the torelon propagators in all cases shows that the continuum dispersion is well satisfied for momenta at any angle to the coordinate axes. In table 3 we give the full set of results for the torelon anisotropies we have measured. There is no noticeable dependence within errors of $`\chi _R`$ on $`L`$ and the dependence on $`\beta `$ is small for the two values of $`\beta `$ used. From the values $`M_T(La_s)`$ we can determine the string tension, $`\sigma `$ in units of $`a_s^2`$. We assume that $`M_T(La_s)=\sigma (La_s)La_s`$ where $`\sigma (La_s)`$ is the string tension modified by finite-size corrections : $`\sigma (La_s)=\sigma +D/(La_s)^2`$. A plot of $`\sigma (La_s)a_s^2`$ versus $`1/L^2`$ is shown in Fig. 5 with a linear fit which has $`\chi ^2/N_{\mathrm{df}}=0.05`$ with $`\sigma a_s^2=0.394(3)`$. In another calculation we have determined the absolute value of $`a_t`$ in $`MeV`$ by computing the splitting $`\mathrm{\Delta }M_{PS}=M_\mathrm{{\rm Y}}(1P)M_\mathrm{{\rm Y}}(1S)`$ for the $`\mathrm{{\rm Y}}`$ meson system using $`O(mv^6)`$ NRQCD . In table 4 we give $`a_t^1,a_s^1,\sigma a_s^2`$, the coefficient $`D`$, and $`\sqrt{\sigma }/\mathrm{\Delta }M_{PS}`$. This latter ratio is experimentally close to unity but in ref. using isotropic $`\beta =6.2`$ UKQCD configurations this ratio was found to be about 1.25 which, from table 4, agrees with our finding except possibly for the $`(\beta =1.7,\chi _0)`$ lattice which has the most coarse lattice spacings. It is generally accepted that the discrepancy is due to quenching and the significant outcome is that our results from the anisotropic lattice agree well with those of a spatially finer isotropic lattice of $`a=0.06`$fm, indicating that we have correctly reproduced the physics expected for this comparison on lattices with $`a_s=0.250.30`$fm. A naive estimate for the coefficient $`D`$ is obtained by calculating the contribution to $`M(L)`$ from the zero-point fluctuations of the torelon regarded as a periodic flux tube. The outcome is $`D=\pi /31.05`$ which is compatible with our fit of $`D1.35(5)`$. The same calculation predicts that there is no $`L`$-independent constant contribution to $`M(L)`$ and this is verified by our fits. The torelon dispersion relation was also studied on the plaquette-tuned parameters. For these simulations, lattices of extent $`(8\times 10)\times 4\times N_t`$ were used, to enable us to investigate a wider range of momenta combinations. At momenta close to the cut-off, we expect the discretization errors to be $`O(a_s^4p^4,\alpha _sa_s^2p^2)`$, thus in order not to contaminate our determination of $`\chi _R`$ with large discretization errors, we first determined the range of momenta over which a good correlated fit to the continuum dispersion relation could be made. The data for different momentum ranges from simulations at $`\beta _P=2.1,\chi _0=6`$ were tested and these results are shown in Table 5. Both the measured anisotropy renormalizations and the torelon rest energies determined are all consistent within statistical precision up to the highest momentum measured. For the subsequent computations of the anisotropy, presented in Table 6, the largest momentum used was the (1,1) data (shown in bold in Table 5). ### 3.2 The Sideways Potential In the sideways potential method , a coarse direction, $`z`$, on the anisotropic lattice is chosen to be the time direction. There are then two types of spacelike direction, coarse and fine. The measurement of the potential between static quarks with a separation lying in the plane of coarse links is compared with the measurement of the potential when the separation lies along the line of fine links. The demand that the two measurements yield the same function of physical distance determines the renormalized anisotropy. Points in the coarse-coarse plane are denoted by $`\stackrel{}{x}=(x,y)`$ and points in the fine direction by $`t`$ where $`x,y,z,t`$ are integers. We measure appropriate spatial Wilson loops $`W_{ss}(\stackrel{}{x},z)`$ and also loops using the fine direction $`W_{ts}(t,z)`$. We define $$V_s(\stackrel{}{x},z)=\mathrm{log}\left(\frac{W_{ss}(\stackrel{}{x},z)}{W_{ss}(\stackrel{}{x},z+1)}\right),V_t(t,z)=\mathrm{log}\left(\frac{W_{ts}(t,z)}{W_{ts}(t,z+1)}\right).$$ (3.2.1) As $`z\mathrm{}`$, $`V_s(\stackrel{}{x},z)V_s(|\stackrel{}{x}|)`$, and $`V_t(t,z)V_t(t)`$ where $`V_s(|\stackrel{}{x}|)`$ and $`V_t(t)`$ are the two versions of the interquark potential. For a physical distance $`r`$ we have $`|\stackrel{}{x}|a_s=ta_t=r`$ . We therefore estimate the renormalized anisotropy $`\chi _R`$ by tuning it so that $`V_s(|\stackrel{}{x}|)=V_t(t/\chi _R)`$, where the right side is evaluated by means of linear interpolation between the values measured at integral $`t`$. It is implicit in the method that there is effective rotational invariance in the $`\stackrel{}{x}`$-plane. We find that if we exclude potentials at the smallest distances, $`|\stackrel{}{x}|=1,\sqrt{2}`$, then the values of $`\chi _R`$ associated with different directions in the $`\stackrel{}{x}`$-plane generally agree within errors. The agreement is particularly good if the links are smeared in an appropriate manner. The results for the renormalized anisotropy, $`\chi _R`$, for both mean-link and plaquette schemes are shown in table 7. An alternative approach to making the comparison is to fit the measured potentials with the forms $`a_sV_s(\stackrel{}{x})=a_sV_0+\sigma a_s^2x+{\displaystyle \frac{e}{x}}`$ $`a_sV_t(t)=a_sV_0+\sigma a_sa_tt+{\displaystyle \frac{a_se}{a_tt}}.`$ (3.2.2) The renormalized anisotropy, $`\chi _R`$, is then determined from the ratio of the coefficients of the linear terms in the two cases. $`\chi _R`$ can in principle be determined from the ratio of the coefficients of the coulombic terms however such an estimate depends on short distance effects and is inherently more sensitive to discretization errors. This approach was tested on an $`8^3\times 48`$ lattice at $`\beta _P=2.3,\chi _0=6`$. Prior to measurement, the lattice is blocked to an $`8^4`$ volume by thinning the time-slices of spatial links and replacing the temporal links with the product of six fine links connecting the appropriate time-slices. The $`t`$ and $`z`$ axes are then interchanged, since as before we want to use a coarse direction for Euclidean decay. Then a standard APE-smearing algorithm is applied to links on the new time slices. These degrees of freedom are then used to measure the potential in the $`(x,y)`$ and $`z`$ axes in the standard fashion. Fig. 6 shows the results of this simulation. The best fits of this data to Eqn. (3.2.2) for a range of static-source separations were computed and the string tension results are shown in Table 8. In Fit C, the coulomb term was poorly resolved and was thus fixed to zero. The use of these different ranges allowed us to investigate the systematic uncertainty and discretization errors in determining $`\chi _R`$ from this method. The three results are consistent within $`2\%`$. The anisotropy measurement from the full range of separations (Fit A, shown in bold in Table 8) is in good agreement with the torelon dispersion result of Table 6. Both the above approaches yielded results reasonably consistent with each other and with the toleron results at the 3% level. Discrepancies can easily be explained by the presence of discretization errors. Because the flux tube generating $`V_s(|\stackrel{}{x}|)`$ has a coarse-fine cross-section and that generating $`V_t(t)`$ a coarse-coarse cross-section we anticipate, as in the case of the toleron (eq.(4.5), that some discretization error in $`\chi _R`$ that is $`O((\alpha _sa_s/R)^2)`$ will remain. Since $`a_s0.3\mathrm{fm}`$, $`R0.7`$ to $`1\mathrm{fm}`$, and $`\alpha _s0.3`$, we expect discretization errors to be around 5%. Because we are concerned to make the long distance behavior consistent in both the fine and coarse directions it is advantageous to use Wilson loops of the largest possible spatial extent. However, in practice, the statistical errors on large Wilson loops grow exponentially with separation. In order to achieve acceptable errors we were restricted to using loops of size 2 or 3 in coarse lattice units. This is to be compared with the torelon, where the practical size is 3 or 4 lattice units. ## 4 Discretization Errors One way of viewing discretization errors is to regard them as arising from the absence of the correction term in the lattice action that yields continuum results. The corresponding term in the action density can be presumed to be a local (redundant) operator. The locality of the operator implies that while there may be a correction to the mass per unit length of the torelon, there is no correction that directly alters the asymptotic proportionality of the torelon mass and its length. The appropriate length scale against which to measure lattice artifacts therefore is the radius $`R`$ of the toleron cross-section or equivalently, $`\mathrm{\Lambda }_{QCD}^1`$ or $`\sqrt{\sigma }`$ where $`\sigma `$ is the string tension. We expect therefore that the $`O(a_s^2)`$ errors in $`\chi _R`$ are proportional to $`(a_s/R)^2`$ . A persuasive plausibility argument for this behavior is as follows. The torelon state of momentum $`𝒑`$ and mass $`M_T`$ is an eigenstate of the transfer matrix $`T`$ with an eigenfunctional labelled by $`𝒑`$ and $`M_T`$ and eigenvalue $`\mathrm{exp}(E(𝒑^2)a_t/\chi _R)`$. Suppose an operator is added to the action which corrects for the $`O(\alpha _sa_s^2)`$ errors. This operator will be local on the scale of the lattice and the effect on $`E(𝒑^2)`$ can be estimated using first-order perturbation theory: $$\delta \frac{a_sE(𝒑^2)}{\chi _R}=\alpha _sf(\overline{𝒑}^2,\overline{M}_T),$$ (4.1) where $`\overline{𝒑}^2=𝒑^2a_s^2`$, $`\overline{M}_T=M_Ta_s`$, and $`f(\overline{𝒑}^2,\overline{M}_T)`$ is the dimensionless matrix element of the added operator which is proportional to $`a_s^2`$ by construction. The question is what scale balances the dimension of $`a_s^2`$? We find $`{\displaystyle \frac{\delta \chi _R}{\chi _R^2}}`$ $`=`$ $`\alpha _s\varphi ^{}(0,\overline{M}_T),`$ (4.2) $`{\displaystyle \frac{\delta \overline{M}_T^2}{2\chi _R}}`$ $`=`$ $`\alpha _s(\varphi (0,\overline{M}_T)\overline{M}_T^2\varphi ^{}(0,\overline{M}_T)),`$ (4.3) where $`\varphi (\overline{𝒑}^2,\overline{M})=a_sE(𝒑^2)f(\overline{𝒑}^2,\overline{M}_T)`$ and $`\varphi ^{}=\varphi /(\overline{𝒑}^2)`$. It must be that the change in the action by a local operator corresponds to a change in the string tension, $`\sigma `$. Hence, since $`M_T=\sigma L`$, it follows that $`\delta \overline{M}_T^2\overline{M}_T^2`$ as this is the only parameter depending on $`L`$, and from eqn. (4.3) this implies that $$\varphi (0,\overline{M}_T)\overline{M}_T^2,\varphi ^{}(0,\overline{M}_T)\overline{M}_T\text{-independent constant}.$$ (4.4) It is conceivable that an accidental cancellation between the two terms in eqn. (4.3) would allow different behavior to be inferred for $`\varphi `$ and $`\varphi ^{}`$ but these results must hold true for all possible local perturbations, not just the particular one that eliminates $`O(\alpha _sa_s^2)`$ errors. We consider this kind of cancellation to be unlikely. Substituting this behavior for $`\varphi ^{}`$ into Eqn. (4.2) we find that $$\frac{\delta \chi _R}{\chi _R^2}C\frac{a_s^2}{R^2},$$ (4.5) where $`R`$ is a typical length scale associated with the torelon which cannot be $`M_T^1`$, for example, the flux tube radius. Alternatively, $`R^1`$ can be taken to be $`\sqrt{\sigma }`$, $`\mathrm{\Lambda }_{QCD}`$. ## 5 Conclusions In this paper we have investigated various methods for measuring the renormalized aspect ratio for pure QCD on an anisotropic lattice. In the main we used a tadpole-improved action with the Landau gauge mean-field definition for the tadpole parameters, but we also included results for the action with tadpoles defined by the mean plaquette. The object of the investigation was to assess the consistency of different methods in order to judge the effectiveness of the improvement scheme. In principle, measurements of the anisotropy from different physical probes should agree close to the Euclidean-symmetric continuum limit. The bare anisotropy $`\chi _0`$ is renormalized by the effect of operators which are irrelevant in the neighborhood of the fixed point controlling the continuum limit. From a renormalization group (RG) view point the location of this fixed point is ambiguous up to redefinitions of the RG transformation used to locate it. This ambiguity is due to redundant operators which have no effect on physical observables of the continuum theory. Consequently, continuum actions with different $`\chi _R`$ can differ only by redundant operators since they must correspond to the same continuum physics. Thus, while $`\chi _R`$ can be changed by tuning $`\chi _0`$, an action with no lattice artifacts must give rotationally invariant physical results and, consequently, any physical method for measuring $`\chi _R`$, such as the ratio of two physical observables, must give the same answer. In as much as this is not the case the differences will give an estimate for the effect of lattice artifacts and the necessity for improvement. The spatial lattice spacing $`a_s`$ was measured in a separate NRQCD simulation by fitting the $`1P1S`$ mass-splitting $`\mathrm{\Delta }M_{PS}`$ for bottomonium. This estimate can be compared with the string tension, deduced from a $`D/L^2`$ extrapolation to $`\mathrm{}`$ of the torelon mass. The coefficient $`D`$ was found to be $`1.38`$ which is in tolerable agreement with $`\pi /3`$ predicted from analysis of flux-tube fluctuations . The ratio $`\sqrt{\sigma }/\mathrm{\Delta }M_{PS}`$ is $`1.2`$ in agreement with earlier NRQCD analyses . The departure from the experimentally observed value of $`1`$ is attributed to quenching. The values for both the mean-link and plaquette tadpoles were found self-consistently. This was very resource intensive in the mean-field case since it required a very accurate gauge fixing to Landau gauge and the self-consistent iteration is in the 2D space of $`(u_s,u_t)`$ which requires additional effort. Also, for given $`\chi _0`$ the tadpoles showed an $`L`$ dependence and hence required an extrapolation to $`L=\mathrm{}`$. This in turn requires accurate data for a good fit. In contrast, the plaquette tadpole was easily found and there was no discernible $`L`$-dependence. The torelon method deduces $`\chi _R`$ from a fully correlated fit to the dispersion relation. Statistical errors are produced by the fit. The sideways potential approach requires some method for matching the coarse and fine potentials and the error analysis is more complex because of both systematic and statistical errors and because the signal rapidly decreases as the loop size increases. Two methods were used to extract $`\chi _R`$. The first compared the loop predictions for the coarse and fine potentials (3.2.1) and deduced $`\chi _R`$ from the rescaling of $`t`$ needed for them to agree. This method was applied to actions with mean-field improvement. Typical loops were of edge length 2-3 in units of the coarse spacing $`a_s`$. There is a clear difference in results compared with the torelon computation of about $`34\%`$. In the second approach the fine direction is first blocked by a scaling of $`\chi _0`$ to give a lattice that is approximately isotropic and then the potential is fitted to a standard form (3.2.2) and $`\chi _R`$ deduced by requiring that the linear slopes agree. The latter method has the advantage that it depends only on the long range structure of the potential and excludes the short range coulomb part. However, it does require large computing resources to extract a reliable signal at large separations, in this case up to 7 lattice units. This technique was applied mainly to actions with plaquette improvement. Although less extensively investigated, the results are consistent with the torelon computation. The observed discrepancies can be easily attributed to residual lattice artifact effects. For the torelon we expect $`\delta \chi _R/\chi _RC\alpha _sa_s^2/R^2`$ which can be sizeable although $`C`$ is unknown. However, it should also be remarked that neither the torelon nor the second sideways potential method include the coulomb part of the static potential. Because of its short-range nature this part will be the most sensitive to the effect of lattice artifacts. In this case agreement of prediction for $`\chi _R`$ is good. This may merely indicate that $`\chi _R`$ is being deduced from the properties of the flux tube in both cases and so they should agree in that artifact effects are the same. However, it is encouraging that different methods based on the long-range properties of the action are consistent and it is worth noting that the torelon flux tube was 3–5 lattice units in length whereas it was necessary to extend the potential method to 7 lattice units. Simulation times are correspondingly reduced for the torelon. Unfortunately, it was not possible to apply the second method to data gathered using the mean-field action because of limitations in resources. We do not find any evidence that the mean-link tuned action is superior to the plaquette tuned action in suppressing lattice artifacts, and it should be noted that the computation required to determine the mean-link tadpoles for the former case is very time consuming. However, there is other evidence that the mean-link action is superior. It is known to give smaller scaling errors in the NRQCD charmonium hyperfine splitting , better hadron mass scaling , and to give a clover coefficient for SW fermions (with the Wilson gauge action) that agrees more closely with the non-perturbatively determined value . Using our configurations we present evidence that it also gives better rotational invariance for the static quark potential. In Fig. 7 we show the deviation of the potential from a fit to the standard form $`V(r)=b+c/r+\sigma r`$. Comparing the mean-link TI data with the plaquette TI data, it is clear that mean-link TI gives smaller deviations from the rotationally invariant fit. In table 9 we present some of the features of Fig. 7 in a quantitative form. As expected, mean-link TI gives better rotational invariance at $`r=\sqrt{3}`$, and it causes the two $`r=3`$ potentials ($`(2,2,1)`$ vs $`(3,0,0)`$) to agree better. As mentioned in the introduction, there are many applications for anisotropic lattices. It would be valuable to measure anisotropies for a wider variety of lattice spacings and bare anisotropies, using the methods we have investigated. It would also be very useful to calculate perturbative formulae for the anisotropy and lattice spacing as a function of the bare parameters of the action, $`\beta `$ and $`\chi _0`$ . Finally, it will be necessary to repeat this tuning process for improved light quark actions. Some of the requisite perturbative calculations have already been performed . ## 6 Acknowledgements This work is supported in part by funds provided by the U.S. Department of Energy (D.O.E.) under cooperative research agreement #DF-FC02-94ER40818. Part of the calculation was performed on the SP-2 at the Cornell Theory Center, which receives funding from Cornell University, New York State, federal agencies, and corporate partners. Calculations were carried out on the Hitachi SR2201 computers at the University of Cambridge High Performance Computing Facility and the Tokyo University Computer Centre.
warning/0003/cond-mat0003444.html
ar5iv
text
# Dynamics of structural models with a long-range interaction : glassy versus non-glassy behavior ## I Introduction The theoretical description of slow dynamics is a crucial point to elucidate the nature of the glass transition in structural glass - forming liquids. One of the commonly used approaches, mode-coupling theory (MCT), was from the very beginning designed for the supercooled simple liquids , i.e. for the non - disordered models (as opposed to the models which contain quenched disorder naturally). Later it has been proven that MC - equations become exact for a number of spin - glass models as well as for the polymeric manifold in a random media (i.e. for the models with quenched disorder) provided that the number of variables components goes to the infinite. Actually applicability of the MC - equations has been substantially extended to the case when the time translation invariance and the fluctuation - dissipation theorem do not hold any more . This striking similarity between the models with and without quenched disorder suggests that the effective disordered potential (e.g., in a supercooled liquid) is in a sense “ self - induced” and the difference between such a “self - induced disorder” and the quenched disorder might not be crucial . In order to provide some insight into self - induced disorder we employed in ref. a Feynman variational principle (VP) for a set of interacting particles. Indeed it was shown that the VP is capable to treat metastable states of the glass - forming system. The main point in ref. was that the partition function representation in terms of functional integrals is twofold : (a) either as an integral over the local density , $`\rho (𝐫)`$, or (b) over the conjugated to $`\rho (𝐫)`$ field $`\psi (𝐫)`$. It has been assumed that the component average free energy , $`\overline{F}`$, (which is only meaningful in the supercooled regime) is equal to the variational free energy $`F_{\mathrm{VP}}`$. There are at least four strong reasons in favor of this (at first sight not obvious) assumption. * The variational free energy $`F_{\mathrm{VP}}`$ is a upper bound for the canonical free energy, i.e., $`F_\mathrm{c}\overline{F}=F_{\mathrm{VP}}`$ , as it should be since $`F_\mathrm{c}=\overline{F}T\mathrm{\Sigma }`$ , where the complexity $`\mathrm{\Sigma }0`$ . * After implementation of VP the initial problem is reduced to a self - consistent random field Ginzburg - Landau model (RFGLM). Then, as was shown previously, the corresponding field $`\psi (𝐫)`$ must be upgraded to a replicated field $`\psi _a(𝐫)`$, where $`a=1,..,n`$ (with the final limit $`n0`$) and the density field $`\rho (𝐫)`$ plays the role of an “external” field. Note that here the density $`\rho (𝐫)`$ is Gaussian due to the use of the VP. Eventually the correlators of $`\psi `$ and $`\rho `$ \- fields can be determined self - consistently. * The resulting replicated partition function for RFGLM has a typical form which may eventually lead to the replica symmetry breaking (RSB), structural glass transition and the “self - induced “ disorder. * Finally, in the case of the long - range interaction the partition function allows the expansion around the saddle point , or mean - field (MF), solution. It is possible to show then that the next to the mean - field approximation and VP merge and both become exact, i. e. $`\overline{F}=F_\mathrm{c}`$ and the glassy phase does not appear. Some evidence for this behavior was deduced from the results for the particles on a $`M`$\- dimensional hypersphere at large dimensions, $`M\mathrm{}`$. The aim of this paper is to face the full dynamical problem for a non - disorded model with a long range interaction. Using the expansion around the saddle point solution we derive the full equation of motion for the time dependent density - density correlator and show that a “glassy” solution does not exist. Conversly, if we add a term describing quenched disorder, by random distribution of the strength of the interaction potential, then the resulting equations of motion for two time density correlation and response functions fall in the same class as MC - equations which have been widely discussed . This means that the “self - induced” disorder is not generic for the pure model with the long-range interaction and vice versa on addition of a quenched disorder the phase space becomes very rugged resulting in slow dynamical processes. The paper is organized as follows. In the next section II we first introduce the theorotical model without quenched disorder . Its dynamics is discussed by using the functional integral technique. The saddle point solution yields the mean field dynamcis. Expansions around the saddle point yield one loop corrections. The Legendre transformation provides the possibilty of the analysis of the full dynamic correlation matrix. In the section III quenched disorder is introduced by a “random bond model” and a Gaussian disorder. The corresponding generating functional (GF) is computed by the self-consistent Hartree approximation, which results in a set of coupled Langevin equations, that are solved in their asymptotic regimes. More details on the calculations are laid out in the corresponding appendices. ## II The model without quenched disorder We start from a simple model system which consists of interacting particles. To do so, let us consider a set of $`N(1)`$ particles in $`d`$ \- dimensional space interacting by a pair potential of the form $`V(r)=\left({\displaystyle \frac{\mu }{N}}\right){\displaystyle \frac{\mathrm{exp}(\kappa r)}{4\pi r^\alpha }}.`$ (1) This is a typical example of a long-range potential with a characteristic length $`\kappa `$ and a coupling constant $`\mu /N`$. The choice of this potential is twofold. It contains a cut off at a $`\kappa ^1`$ and allows thus to control the range of the interaction. Moreover at small scales $`(r<\kappa ^1)`$ it consists of a typical power law decay with long range character, if $`0<\alpha <2`$. Therefore the so chosen potential allows to keep control on range and nature of the interaction, which will become essential below. To ensure extensivity of the total interaction energy we require that the integral $`d^dV(𝐫)=𝒪(N^0)`$, i.e. it does not depend on the number of particles $`N`$. As a result we have $`\kappa N^{1/(d\alpha )}`$. The intermolecular potential (1) has the form of the generalized Kac potential $`V(𝐫)=\kappa ^df(\kappa 𝐫),`$ (2) which has been used for the rigorous treatment of the van-der-Waals theory . In order to provide conditions for the expansion around a saddle point, carried out later on (see below), we should require that the length $`\kappa ^1`$ must be larger compared to the characteristic size of the system, which scales naturally as $`N^{1/d}`$) at $`N\mathrm{}`$. As a consequence we find the limits for the range parameter $`\alpha `$ $`0<\alpha <d.`$ (3) Below we shall restrict our considerations to the case: $`d=3,\alpha =1`$, and the strength of the interaction $`\mu >0`$ (pure repulsion) without loss of generalization in the main statements we are going to predict. Then the Fourier transformation of the potential (1) takes especially simple form $`V(𝐤)=\left({\displaystyle \frac{\mu }{N}}\right){\displaystyle \frac{1}{k^2+\kappa ^2}},`$ (4) which allows accurate analytic calculations. In the limit $`N\mathrm{}`$ we have thus $`\kappa ^2N^1`$, but the relevant minimum wave vector is $`k_{\mathrm{min}}^2N^{2/3}`$ and thus $`\kappa ^2`$ can be actually neglected under the integration over the whole $`k`$ \- space. As a result we arrive formally at a one - component plasma model (OCP) where the electroneutrality is implicitly provided by a neutralizing background. ### A The generating functional method In the following we set up the relevant equations of motion for the model system. We restrict ourselves to the Langevin dynamics, which can be comfortably formulated in terms of dynamic functionals which allows the systematic $`1/N`$ \- expansion treatment. The Langevin dynamics of $`N`$ particles interacting via the potential (1) (at $`d=3,\alpha =1`$ and $`\mu >0`$) is described by the equation of motion $`m_0{\displaystyle \frac{^2}{t^2}}𝐫^{(p)}(t)+\gamma _0{\displaystyle \frac{}{t}}𝐫^{(p)}(t){\displaystyle \frac{\mu }{N}}{\displaystyle \underset{m=1}{\overset{N}{}}}v(𝐫^{(p)}𝐫^{(m)})=𝐟^{(p)}(t),`$ (5) where $`m_0`$ and $`\gamma _0`$ are the mass and the friction coefficient respectively, $`p=1,2,\mathrm{},N`$ and $`v(r;\kappa )=\mathrm{exp}(\kappa r)/4\pi r`$. The random force in eq.(5) is Gaussian with $`f_i^{(p)}(t)=0`$ and the correlator $`f_i^{(p)}(t)f_j^{(n)}(t^{})=2T\gamma _0\delta _{pm}\delta _{ij}\delta (tt^{}),`$ (6) where from now on we work in units where the Boltzmann constant $`\mathrm{k}_\mathrm{B}=1`$. As was mentioned, it is more convenient to reformulate the Langevin problem (5)-(6) by using the celebrated Martin-Siggia-Rose generating functional (GF) method . The method was first applied for the $`\varphi ^4`$ \- model with the long-range interaction in and for the polymer melt dynamics in . Despite the fact that the Langevin equation (5) is of the second order it is possible to show that the Jacobian which appear under transformation to the functional variables, still equal to one (see Appendix in ). After using this technique for the problem (5) - (6), GF takes the form $`Z\{\mathrm{}\}={\displaystyle \underset{p=1}{\overset{N}{}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\left\{\underset{p=1}{\overset{N}{}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]+𝑑t\underset{p=1}{\overset{N}{}}\underset{m=1}{\overset{N}{}}\frac{\mu }{N}i\widehat{r}_j^{(p)}(t)_j^{(p)}v\left(𝐫^{(p)}𝐫^{(m)}\right)\right\}},`$ (7) where the action of the free system $`A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]={\displaystyle 𝑑t\left\{T\gamma _0\left[i\widehat{r}_j^{(p)}(t)\right]^2+i\widehat{r}_j^{(p)}(t)\left[m_0\frac{^2}{t^2}r_j^{(p)}(t)+\gamma _0\frac{}{t}r_j^{(p)}(t)\right]\right\}}.`$ (8) In the following we are going to transform this functional to collective density variables. By using the transformations to the mass density $`\rho (𝐫)={\displaystyle \underset{p=1}{\overset{N}{}}}\delta (𝐫𝐫^{(p)}(t))`$ (9) and the longitudinal projection of the response field density $`\pi (𝐫)={\displaystyle \underset{p=1}{\overset{N}{}}}i\widehat{r}_i^{(p)}(t)_i\delta (𝐫𝐫^{(p)}(t))`$ (10) for the GF one gets $`Z\left\{\chi _\alpha \right\}={\displaystyle \underset{\alpha =0}{\overset{1}{}}D\rho _\alpha (1)\mathrm{exp}\left\{W\{\rho _\alpha \}\frac{1}{2}d1d2\rho _\alpha (1)U_{\alpha \beta }(1,2)\rho _\beta (2)+d1\rho _\alpha (1)\chi _\alpha (1)\right\}},`$ (11) where the summation over the repeated Greek indices is implied. In eq.(11) we have introduced 2-dimensional field $`\rho _\alpha (1)\left({\displaystyle \genfrac{}{}{0pt}{}{\rho (1)}{\pi (1)}}\right),`$ (12) where $`\alpha =0,1`$ and $`1(𝐫,t)`$. The “entropy” of the free system is given as usual by $`W\{\rho ,\pi \}`$ $`=`$ $`\mathrm{log}{\displaystyle \underset{p=1}{\overset{N}{}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\left\{\underset{p=1}{\overset{N}{}}A_0\{𝐫^{(p)},\widehat{𝐫}^{(p)}\}\right\}}`$ (13) $`\times `$ $`\delta \left[\rho (𝐫,t){\displaystyle \underset{p=1}{\overset{N}{}}}\delta (𝐫𝐫^{(p)}(t))\right]`$ (14) $`\times `$ $`\delta \left[\pi (𝐫,t){\displaystyle \underset{p=1}{\overset{N}{}}}i\widehat{r}_j^{(p)}(t)_j\delta (𝐫𝐫^{(p)}(t))\right],`$ (15) and $`U_{\alpha \beta }`$ is the 2$`\times `$ 2 - interaction matrix $`U_{\alpha \beta }(1,2)=\left({\displaystyle \genfrac{}{}{0pt}{}{0}{V(|𝐫_1𝐫_2|)}}{\displaystyle \genfrac{}{}{0pt}{}{V(|𝐫_1𝐫_2|)}{0}}\right)`$ (16) and $`\chi _\alpha (1)`$ is a source field. An alternative valuable representation of GF can be obtained through the “functional Fourier transformation” $`\mathrm{exp}\left\{F\{\psi _\alpha \}\right\}={\displaystyle D\rho _\alpha (1)\mathrm{exp}\left\{W\{\rho _\alpha \}id1\rho _\alpha (1)\psi _\alpha (1)\right\}}`$ (17) and its inversion $`\mathrm{exp}\left\{W\{\rho _\alpha \}\right\}={\displaystyle D\psi _\alpha (1)\mathrm{exp}\left\{F\{\psi _\alpha \}+id1\rho _\alpha (1)\psi _\alpha (1)\right\}}.`$ (18) The substitution of eq.(15) into eq.(17) leads to the explicit expression for the free-system GF $`\mathrm{exp}\left\{F\{\psi _\alpha \}\right\}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{p=1}{\overset{N}{}}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]`$ (19) $``$ $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dt\psi \left(𝐫^{(p)}\right)+i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dti\widehat{r}_j^{(p)}(t)_j\varphi (𝐫)|_{𝐫=𝐫^{(p)}(t)}\},`$ (20) where $`\psi (1)`$ and $`\varphi (1)`$ are components of the column - variable $`\psi _\alpha (1)\left({\displaystyle \genfrac{}{}{0pt}{}{\psi (1)}{\varphi (1)}}\right).`$ (21) By making use (18) in (11) and after functional integration over $`\rho _\alpha (1)`$ one gets $`Z\{\chi _\alpha ,\lambda _\alpha \}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{\alpha =0}{\overset{1}{}}}D\psi _\alpha (1)\mathrm{exp}\{F\{\psi _\alpha \}+`$ (22) $`{\displaystyle \frac{1}{2}}{\displaystyle }d1d2[i\psi _\alpha (1)`$ $`+`$ $`\chi _\alpha (1)][U^1]_{\alpha \beta }(1,2)[i\psi _\beta (2)+\chi _\beta (2)]+{\displaystyle }d1\psi _\alpha (1)\lambda _\alpha (1)\},`$ (23) where we have also add a source field $`\lambda _\alpha (1)`$ conjugated to $`\psi _\alpha (1)`$. As a result eqs.(11) and (23) provide two equivalent representations of GF. For the purpose of expansion around the saddle point we use representation (23) at $`\lambda _\alpha (1)=0`$ which after the transformation $`\psi _\alpha \psi _\alpha +i\chi _\alpha `$, yields $`Z\left\{\chi _\alpha \right\}={\displaystyle \underset{\alpha =0}{\overset{1}{}}D\psi _\alpha (1)\mathrm{exp}\left\{NA[\psi _\alpha ;\chi _\alpha ]\right\}},`$ (24) which is appropriate for a saddle point integration, since the particle number $`N`$ is large. The action hereby is given as $`A[\psi _\alpha ;\chi _\alpha ]`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑td^3𝐫d^3𝐫^{}\psi _\alpha (𝐫,t)\left[v^1\right]_{\alpha \beta }(𝐫𝐫^{};\kappa )\psi _\beta (𝐫^{},t)}{\displaystyle \frac{1}{N}}\mathrm{log}{\displaystyle \underset{p=1}{\overset{N}{}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)}`$ (25) $`\times `$ $`\mathrm{exp}\left\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle 𝑑tr_\alpha ^{(p)}(t)\left[\psi _\alpha \left(𝐫^{(p)}(t)\right)+i\chi _\alpha \left(𝐫^{(p)}(t)\right)\right]}\right\},`$ (26) and the interaction matrix $`v_{\alpha \beta }(𝐫;\kappa )=\left({\displaystyle \genfrac{}{}{0pt}{}{0}{1}}{\displaystyle \genfrac{}{}{0pt}{}{1}{0}}\right){\displaystyle \frac{\mathrm{exp}(\kappa r)}{4\pi r}}.`$ (27) Recall that the relation $`\kappa N^{1/2}`$ is necessary for the validity of the saddle point integration. Moreover we have defined the column-vector $`r_\alpha ^{(p)}(t)=\left({\displaystyle \genfrac{}{}{0pt}{}{1}{i\widehat{r}_j^{(p)}(t)𝑑\tau \frac{\delta }{\delta r_j^{(p)}(\tau )}}}\right)`$ (28) for convenience. ### B The saddle point solution and expansion around the SP Minimization of $`A[\psi _\alpha ;\chi _\alpha ]`$ with respect to $`\psi _\alpha (1)`$ leads to the SP - equations for the mean fields $`\overline{\psi _\alpha }(1)`$ $`\overline{\psi _\alpha }(𝐫,t)={\displaystyle \frac{i\mu }{N}}{\displaystyle d^3𝐫^{}v_{\alpha \beta }(𝐫𝐫^{})\rho _\beta (𝐫^{},t)_{\mathrm{SP}}},`$ (29) where the average $`\mathrm{}_{SP}`$ is calculated by using the cumulant GF $`P_{\mathrm{SP}}\{\overline{\psi _\alpha }+i\chi _\alpha \}{\displaystyle \frac{1}{N}}\mathrm{log}{\displaystyle }{\displaystyle \underset{p=1}{\overset{N}{}}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]`$ (30) $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dtr_\alpha ^{(p)}(t)[\overline{\psi _\alpha }\left(𝐫^{(p)}(t)\right)+i\chi _\alpha \left(𝐫^{(p)}(t)\right)]\}.`$ (31) The correlation matrix in the random phase approximation (RPA) is defined in such a way $`S_{\alpha \beta }(1,2)=\underset{\overline{\psi _\alpha }+i\chi _\alpha 0}{lim}\left[{\displaystyle \frac{\delta \rho _\alpha (1)_{\mathrm{SP}}}{N\delta \chi _\beta (2)}}\right].`$ (32) After linearization of eq.(31) with respect to $`\overline{\psi _\alpha }+i\chi _\alpha `$ the 2$`\times `$2 - RPA correlation matrix is easily found to coincide with the well known form $`S_{\alpha \beta }(1,2)=\left\{\left[\widehat{F}^1+\mu \widehat{v}\right]^1\right\}_{\alpha \beta }(1,2),`$ (33) where $`\widehat{v}`$ is the interaction matrix (27) and $`F_{\alpha \beta }`$ is the correlation matrix for the free system $`F_{\alpha \beta }(1,2)=\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)_0/N`$ has the form $`F_{\alpha \beta }(1,2)=\left({\displaystyle \genfrac{}{}{0pt}{}{F_{00}(1,2)}{F_{10}(1,2)}}{\displaystyle \genfrac{}{}{0pt}{}{F_{01}(1,2)}{0}}\right).`$ (34) In eq.(34) $`F_{01}(1,2)`$ and $`F_{10}(1,2)`$ are response functions whereas $`F_{00}(1,2)`$ stands for the correlation function. The relation between them is given by the fluctuation dissipation theorem (FDT) which in $`(𝐤,t)`$\- representation has the form $`\beta {\displaystyle \frac{}{t}}F_{00}(𝐤,t)=F_{01}(𝐤,t)F_{10}(𝐤,t).`$ (35) It is easy to check that in this case the FDT for the RPA type correlation matrix (33) also holds $`\beta {\displaystyle \frac{}{t}}S_{00}(𝐤,t)=S_{01}(𝐤,t)S_{10}(𝐤,t),`$ (36) where $`\beta =1/T`$ is the inverse temperature. The corresponding elements of the RPA - matrix (33) are of an especially simple form in the Fourier - $`(𝐤,\omega )`$ \- representation, namely $`S_{00}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{F_{00}(𝐤,\omega )}{\left[1+v(k)F_{10}(𝐤,\omega )\right]\left[1+v(k)F_{01}(𝐤,\omega )\right]}}`$ (37) $`S_{01}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{F_{01}(𝐤,\omega )}{1+v(k)F_{01}(𝐤,\omega )}}`$ (38) $`S_{10}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{F_{10}(𝐤,\omega )}{1+v(k)F_{10}(𝐤,\omega )}}.`$ (39) It turns out interesting to recover the wellknown form in the static limit, where we have $`S_{01}(𝐤,\omega 0)=\beta S_{\mathrm{st}}(𝐤)=\left[\left(\beta F_{\mathrm{st}}\right)^1+\mu k^2\right]`$ and for the correlator $`S_{\mathrm{RPA}}(𝐤)=S_{\mathrm{st}}(𝐤)/\rho _0`$ one gets $`S_{\mathrm{RPA}}(𝐤)={\displaystyle \frac{1}{1+\frac{\beta \mu \rho _0}{k^2}}},`$ (40) where we have used $`F_{\mathrm{st}}=\rho _0`$ . This expression is completely equivalent to the correlator for the OCP-model (see eq.(10.1.7) in ) with the direct correlation function $`c(𝐤)=\mu \beta /k^2`$ and the Debye wavenumber $`k_\mathrm{D}=\left(\beta \mu \rho _0\right)^{1/2}`$. Now let us expand the action (26) around SP-solution (29) up to the second order with respect to the fluctuations $`\psi _\alpha (1)\overline{\psi _\alpha }(1)`$. After the functional integration we arrive at the following result for the GF $`P\left\{\chi _\alpha \right\}`$ $``$ $`{\displaystyle \frac{1}{N}}\mathrm{log}Z\left\{\chi _\alpha \right\}`$ (41) $`=`$ $`P_{\mathrm{SP}}\left\{\overline{\psi _\alpha }+\chi _\alpha \right\}{\displaystyle \frac{1}{2N}}\mathrm{Tr}\left[\mathrm{log}T_{\alpha \beta }(1,2)\right],`$ (42) where $`T_{\alpha \beta }(1,2)`$ is the inverse matrix of the effective interactions $`T_{\alpha \beta }(1,2)={\displaystyle \frac{1}{\mu }}\left[v^1\right]_{\alpha \beta }(1,2)+{\displaystyle \frac{1}{N}}\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)_{\mathrm{SP}}.`$ (43) In eqs.(42) - (43) we deliberately keep external field $`\chi _\alpha (1)`$ nonzero because it to be used in the next subsection for the Legendre transformation. ### C The Legendre transformation The functional Legendre transformation is a general way to provide the Dyson equation for the full correlation matrix $`G_{\alpha \beta }(1,2)`$ . In doing so the irreducible GF, $`\mathrm{\Gamma }\left\{\rho _\alpha (1)\right\}`$ , is defined by the identity $`\mathrm{\Gamma }\left\{\rho _\alpha (1)\right\}+P\left\{\chi _\alpha (1)\right\}={\displaystyle d1\rho _\alpha (1)\chi _\alpha (1)}.`$ (44) By doing functional differentiation of (44) one gets $`\chi _\alpha (1)={\displaystyle \frac{\delta \mathrm{\Gamma }\left\{\rho _\alpha (1)\right\}}{\delta \rho _\alpha (1)}}`$ (45) and $`\left[G^1\right]_{\alpha \beta }(1,2)={\displaystyle \frac{\delta ^2\mathrm{\Gamma }\left\{\rho _\alpha (1)\right\}}{\delta \rho _\alpha (1)\delta \rho _\beta (2)}}.`$ (46) Taking into account the result in eq.(42) we find the following result for GF $`\mathrm{\Gamma }\left\{\rho _\alpha (1)\right\}=\mathrm{\Gamma }_{\mathrm{SP}}\left\{\rho _\alpha (1)\right\}+{\displaystyle \frac{1}{2N}}\mathrm{Tr}\left[\mathrm{log}T_{\alpha \beta }(1,2)\right],`$ (47) where $`\mathrm{\Gamma }_{\mathrm{SP}}\left\{\rho _\alpha (1)\right\}=P_{\mathrm{SP}}\left\{\chi _\alpha \right\}+{\displaystyle d1\rho _\alpha (1)\chi _\alpha (1)}.`$ (48) In eq.(47) one should consider $`\chi _\alpha (1)`$ as a functional of $`\rho _\alpha (1)`$ given by eq.(45). Double differentiation of eq.(47) leads to an equation of the Dyson form $`\left[G^1\right]_{\alpha \beta }(1,2)=\left[S^1\right]_{\alpha \beta }(1,2)\mathrm{\Sigma }_{\alpha \beta }(1,2),`$ (49) where the RPA - correlation matrix, $`S_{\alpha \beta }(1,2)`$, is defined by eqs.(37) - (39) and the “self - energy” functional $`\mathrm{\Sigma }_{\alpha \beta }(1,2)`$ has the form $`\mathrm{\Sigma }_{\alpha \beta }(1,2)={\displaystyle \frac{1}{2N}}\mathrm{Tr}\left\{{\displaystyle \frac{\delta ^2}{\delta \rho _\alpha (1)\delta \rho _\beta (2)}}\mathrm{log}T_{\gamma \delta }(3,4)\right\}_{\chi _\alpha =0}.`$ (50) In eq.(50) the “trace” is taken over the variables 3,4 and indices $`\gamma ,\delta `$. The explicit differentiation in eq.(50) leads to the result $`\mathrm{\Sigma }_{\alpha \beta }(1,2)={\displaystyle \frac{1}{2N}}\mathrm{Tr}\left\{\widehat{T}^1{\displaystyle \frac{\delta ^2\widehat{T}}{\delta \rho _\alpha (1)\delta \rho _\beta (2)}}\right\}_{\chi _\alpha =0},`$ (51) where $`\widehat{T}`$ is a short hand notation of the matrix $`T_{\gamma \delta }(3,4)`$ and we have took into account that $`\delta T_{\alpha \beta }(1,2)/\delta \chi _\gamma (3)=\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)\mathrm{\Delta }\rho _\gamma (3)_{\mathrm{SP}}/N=0`$ at $`\chi _\alpha =0`$ because the fluctuations are Gaussian. Further calculation yields $`{\displaystyle \frac{\delta ^2T_{\gamma \delta }(3,4)}{\delta \rho _\alpha (1)\delta \rho _\beta (2)}}={\displaystyle d5d6\frac{1}{N}\mathrm{\Delta }\rho _\gamma (3)\mathrm{\Delta }\rho _\delta (4)\mathrm{\Delta }\rho _\omega (5)\mathrm{\Delta }\rho _\chi (6)_{\mathrm{SP}}R_{\omega \beta }(5,2)R_{\chi \alpha }(6,1)},`$ (52) where $`R_{\alpha \beta }(1,2)={\displaystyle \frac{\delta \vartheta _\alpha (1)}{\delta \rho _\beta (2)}}`$ (53) and the full mean field $`\vartheta _\alpha (1)=i\overline{\psi _\alpha }(1)+\chi _\alpha (1).`$ (54) The expression for $`R_{\alpha \beta }(1,2)`$ can be easily found by differentiation of eq.(54) with respect to $`\rho _\beta (2)`$. Taking into account eqs.(29), (45) and (46) at $`\chi _\alpha 0`$ one gets $`R_{\alpha \beta }(1,2)=\left[G^1\right]_{\alpha \beta }(1,2)\mu {\displaystyle d4d3v_{\alpha \omega }(1,4)S_{\omega \gamma }(4,3)R_{\gamma \beta }(3,2)}`$ (55) or finally $`R_{\alpha \beta }(1,2)={\displaystyle d3\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{S}\right]^1\right\}_{\alpha \gamma }(1,3)\left[\widehat{G}^1\right]_{\gamma \beta }(3,2)},`$ (56) where the hatted variables stands for the corresponding 2$`\times `$2 matrices. Substitution eqs.(56) and (52) in eq.(51) yields $`\mathrm{\Sigma }_{\alpha \beta }(1,2)={\displaystyle d3d4K_{\gamma \delta }(3,4)\left[G^1\right]_{\gamma \alpha }(3,1)\left[G^1\right]_{\delta \beta }(4,2)},`$ (57) where 2$`\times `$2 vertex - matrix has the form $`K_{\alpha \beta }(1,2)`$ $`=`$ $`{\displaystyle d3d4d5d6\left\{\left[(\mu \widehat{v})^1+\widehat{S}\right]^1\right\}_{\delta \gamma }(4,3)S_{\gamma \delta \omega \chi }^{(4)}(3,4,5,6)}`$ (58) $`\times `$ $`\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{S}\right]^1\right\}_{\omega \alpha }(5,1)\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{S}\right]^1\right\}_{\chi \beta }(6,2).`$ (59) In eq.(59) $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ is the 4 - point (response) correlator matrix in RPA $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)={\displaystyle \frac{1}{N^2}}\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)\mathrm{\Delta }\rho _\gamma (3)\mathrm{\Delta }\rho _\delta (4)_{\mathrm{SP}}.`$ (60) The explicit calculation of $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ is implemented in Appendix B. The vertex - matrix can be seen as a one-loop diagram (see Fig. 1). The higher loops contributions which include generally speaking 2m - point correlators , $`S_{\alpha \beta \mathrm{}\gamma }^{(2m)}(1,2,\mathrm{},2m)`$, can be also considered, however the “self - energy” has still the same convolution structure : $`\widehat{\mathrm{\Sigma }}=\widehat{G}^1\stackrel{~}{K}\widehat{G}^1`$. Here the vertex - matrix $`\stackrel{~}{K}_{\alpha \beta }(1,2)`$ is calculated in RPA only . That is why these contributions basically do not change our results. As a result in the $`(𝐤,\omega )`$\- representation the Dyson equation (49) with the “self - energy” functional (57) reduces to a quadratic one $`G_{\alpha \gamma }(𝐤,\omega )\left[S^1\right]_{\gamma \delta }(𝐤,\omega )G_{\delta \beta }(𝐤,\omega )G_{\alpha \beta }(𝐤,\omega )+K_{\alpha \beta }(𝐤,\omega )=0.`$ (61) The coefficients of the eq.(61) trace the problem back to the free system dynamics which is embodied in the correlation matrices $`F_{\alpha \beta }(1,2)`$ and $`F_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$. It is not surprising then that a specification of the model for the free system dynamics is necessary, before going to the investigation of eq.(61). ### D Analysis of the equation for the full correlation matrix As we have mentioned the explicit solution of eq.(61) needs the specification of the free system dynamics. Two simple models are most amenable for the theoretical treatment : free diffusion model (FDM) and the relaxation time approximation model (RTAM ) . The latter provide more reasonable dynamical information also for short time intervals, $`\mathrm{\Delta }t<m_0/\gamma _0`$ , where the FDM completely failed (e.g., the sum rule does not hold). It turns out that upon calculation of the trace in eq.(59) the integral is ultraviolet \- divergent for FDM and only RTAM leads to the finite result. The matrix elements for RTAM has the form $`F_{00}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{2F_{\mathrm{st}}k^2D}{\omega ^2+(k^2D\omega ^2\tau _0)^2}}`$ (62) $`F_{01}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{\beta F_{\mathrm{st}}k^2D}{i\omega +k^2D\omega ^2\tau _0}}`$ (63) $`F_{10}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{\beta F_{\mathrm{st}}k^2D}{i\omega +k^2D\omega ^2\tau _0}},`$ (64) where we introduced the diffusion coefficient $`D=T/\gamma _0`$, the characteristic time scale $`\tau _0=m_0/\gamma _0`$, and $`F_{\mathrm{st}}=\rho _0`$ for the overall density. At $`\tau _0=0`$ we return to FDM. In the case of RTAM the solution of eq.(61) for the full correlation matrix reads $`G_{01}(𝐤,\omega )`$ $`=`$ $`{\displaystyle \frac{1+\sqrt{14\left[i\omega \tau _\mathrm{c}\omega ^2\tau _0\tau _\mathrm{c}+\chi _{st}^1\right]K_{01}(𝐤,\omega )}}{2\left[i\omega \tau _\mathrm{c}\omega ^2\tau _0\tau _\mathrm{c}+\chi _{st}^1(k)\right]}}`$ (65) $`G_{10}(𝐤,\omega )`$ $`=`$ $`G_{01}(𝐤,\omega )`$ (66) and $`G_{00}(𝐤,\omega )={\displaystyle \frac{\frac{\tau _\mathrm{c}}{2\beta }\left|\frac{1+\sqrt{14\left[i\omega \tau _\mathrm{c}\omega ^2\tau _0\tau _\mathrm{c}+\chi _{st}^1(𝐤)\right]K_{01}(𝐤,\omega )}}{i\omega \tau _\mathrm{c}\omega ^2\tau _0\tau _\mathrm{c}+\chi _{st}^1(𝐤)}\right|^2K_{00}(𝐤,\omega )}{\mathrm{Re}\left\{\sqrt{14\left[i\omega \tau _\mathrm{c}\omega ^2\tau _0\tau _\mathrm{c}+\chi _{st}^1(𝐤)\right]K_{01}(𝐤,\omega )}\right\}}}.`$ (67) The explicit calculation of the matrix $`K_{\alpha \beta }(𝐤,\omega )`$ (see eq.(59) is given in the Appendix C. The overall behavior of the correlation function $`G_{00}(𝐤,\omega )`$ according eq. (67) is shown in Fig. 2 (at $`\mu =10,\beta =0.1,\rho _0=1\text{and}\tau _0=0.1`$). It can be seen clearly that there no singularity appears at $`\omega 0`$. The singularity, however, might be responsible for a glass transition. Instead, the low frequency limit of $`G_{00}(𝐤,\omega )`$ slowly changed with control parameters (which is not shown in Fig.2). That means that for the non - disordered model with a general repulsive long ranged potential (1) the glass transition is not generic. This very important conclusion suggest that for the model with a long range interaction the phase space is too smooth to show a glass transition. In order to obtain a glass like transition a competing interactions or a quenched disorder should be added. This leads to glassy dynamics, as we will show in the next section. It is interesting to note, that for the generalized Kac potential eq. (2), where $`f(𝐫)`$ and its Fourier transformation are positive definite functions, MCT - memory kernel vanishes at $`\kappa 0`$. The corresponding argumentation is relegated to the Appendix D. The explanation for this result lies in the fact that the “cage effect”, which is a cornerstone of MCT, is missing in the MF - limit. The “glass transition” which has been studied in ref. for the particles interacted via the Kac potential (2) has a completely different nature. In ref. the function $`f(𝐫)`$ has a step form so that its Fourier transform $`f(𝐤)`$ is negative at some value of $`𝐤`$ . As a result the system becomes unstable and a nonuniform configuration where the particles are grouped into “clumps” shows up. It was found that the slow dynamics of the MF - model is associated with these clumps and does not touch a single particle motion. Obviously it is different from the conventional glass transition . ## III The structural model with a competing quenched interactions ### A Specification of the model In the previous sections we have shown in detail that in the absence of disorder the dynamical spectrum changed monotonically with a control parameter and no glassy dynamics can be seen. The natural question which arises now is: How will the introduction of competing interactions and / or quenched disorder affect the dynamics of the system discussed above? To provide an answer to this question we will use already existing models of heteropolymers and their disordered two body interaction . The use of these models and techniques are here natural, since the behavior of heteropolymers is well discussed in the literature. In principle, two practical possibilities exist. * The strength of the two- body interaction , $`\mu `$, in eq.(1) is now a random function of all pairs of the interacted particles, $`\mu _{pm}`$ (“random - bond model”). * Each particle carry a single “charge” $`\sigma _p`$, so that $`\mu _{pm}=\xi _0+b\sigma _p\sigma _m`$ and each $`\sigma _p`$ is randomly distributed (“random sequence model”). It turns out to be sufficient for the purpose of this paper to restrict ourself only with the “random - bond model” , where $`\mu _{pm}`$ does not depend from the choice of pairs and has a Gaussian distribution $`P\{\mu _{pm}\}\mathrm{exp}\left\{{\displaystyle \frac{(\mu _{pm}\xi _0)^2}{2\chi ^2}}\right\}.`$ (68) The competing long range interactions frustrates the system of particles and the question is whether a glass transition exist or not. Normally frustration and frozen disorder is enough for the existence of glassy phases. Here the problem is more complicated, since the long range nature of the interaction may provide opposite effects. The averaging over the quenched disorder in eq.(7) (after the substitution $`\mu \mu _{pm})`$ can be carried out just in the same way as in ref.. Similarly, typical two-time dependent terms immediately appear. They are also bilinear with respect to the forces of interaction $`_jv(𝐫)`$ . In order to rationalize these terms it is convenient to introduce (besides the mass density (9) and the response field density (10)) the following collective variables $`Q_0(𝐫,t;𝐫^{},t^{})`$ $`=`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}\delta (𝐫𝐫^p(t))\delta (𝐫^{}𝐫_{}^{}{}_{}{}^{p}(t^{}))`$ (69) $`Q_1(𝐫,t;𝐫^{},t^{})`$ $`=`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}i\widehat{r}_j(t)_j\delta (𝐫𝐫^p(t))i\widehat{r}_l(t^{})_l\delta (𝐫^{}𝐫_{}^{}{}_{}{}^{p}(t^{}))`$ (70) $`Q_2(𝐫,t;𝐫^{},t^{})`$ $`=`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}i\widehat{r}_j(t)_j\delta (𝐫𝐫^p(t))\delta (𝐫^{}𝐫_{}^{}{}_{}{}^{p}(t^{}))`$ (71) $`Q_3(𝐫,t;𝐫^{},t^{})`$ $`=`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}i\widehat{r}_j(t^{})_j\delta (𝐫^{}𝐫^p(t^{}))\delta (𝐫𝐫^p(t)).`$ (72) After the introduction of the 4-dimensional column - fields $`Q_\mathrm{a}(1;1^{})=\left[\begin{array}{c}Q_0(1;1^{})\\ Q_1(1;1^{})\\ Q_2(1;1^{})\\ Q_3(1;1^{})\end{array}\right],`$ (77) where $`a=1,2,3,4`$and 4$`\times `$4 - matrix $`\mathrm{\Gamma }_{\mathrm{ab}}(1,2,3,4)=\left[\begin{array}{cccc}0& v(1,3)v(2,4)& 0& 0\\ v(1,3)v(2,4)& 0& 0& 0\\ 0& 0& 0& v(1,4)v(3,2)\\ 0& 0& v(1,4)v(3,2)& 0\end{array}\right]`$ (82) the whole expression for GF takes the form $`Z_{\mathrm{av}}\{\chi _\alpha ,H_a\}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{\alpha =0}{\overset{1}{}}}{\displaystyle \underset{a=0}{\overset{3}{}}}D\rho _\alpha (1)DQ_\mathrm{a}(1;1^{})\mathrm{exp}\{\stackrel{~}{W}\{\rho _\alpha (1);Q_\mathrm{a}(1;1^{})\}`$ (83) $``$ $`{\displaystyle \frac{\xi _0}{2N}}{\displaystyle d1d2\rho _\alpha (1)U_{\alpha \beta }(1,2)\rho _\beta (2)}+{\displaystyle d1\rho _\alpha (1)\chi _\alpha (1)}`$ (84) $``$ $`{\displaystyle \frac{\chi ^2}{4N^2}}{\displaystyle }d1d2d3d4Q_\mathrm{a}(1;2)\mathrm{\Gamma }_{\mathrm{ab}}(1,2,3,4)Q_\mathrm{b}(3;4)+{\displaystyle }d1d2Q_\mathrm{a}(1;2)H_a(1;2)\},`$ (85) where the entropy is given by $`\stackrel{~}{W}\{\rho _\alpha (1);Q_\mathrm{a}(1;1^{})\}`$ $`=`$ $`\mathrm{log}{\displaystyle \underset{p=1}{\overset{N}{}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\left\{A_0\{𝐫^{(p)},\widehat{𝐫}^{(p)}\}\right\}}`$ (86) $`\times `$ $`{\displaystyle \underset{\alpha =0}{\overset{1}{}}}\delta \left[\rho _\alpha (1){\displaystyle \underset{p=1}{\overset{N}{}}}r_\alpha ^{(p)}(1)\delta (𝐫_1𝐫^{(p)}(t))\right]`$ (87) $`\times `$ $`{\displaystyle \underset{\mathrm{a}=1}{\overset{4}{}}}\delta \left[Q_\mathrm{a}(1;2){\displaystyle \underset{p=1}{\overset{N}{}}}p_\mathrm{a}^{(p)}(1;2)\delta (𝐫_1𝐫^{(p)}(t_1))\delta (𝐫_2𝐫^{(p)}(t_2))\right].`$ (88) We had used the column - operators $`r_\alpha ^{(p)}(1)=\left(\begin{array}{c}1\\ i\widehat{r}_j^{(p)}(t_1)_{j,1}\end{array}\right)p_\mathrm{a}^{(p)}(1;2)=\left(\begin{array}{c}1\\ i\widehat{r}_j^{(p)}(t_1)_{j,1}i\widehat{r}_l^{(p)}(t_2)_{l,2}\\ i\widehat{r}_j^{(p)}(t_1)_{j,1}\\ i\widehat{r}_j^{(p)}(t_2)_{j,2}\end{array}\right)`$ (95) and the external field, $`H_\mathrm{a}(1;2)`$, conjugated to $`Q_\mathrm{a}(1;2)`$, has been introduced also. The two- point collective fields (72) have a meaning of the dynamical “overlaps”. It is a dynamical generalization of the Parisi “overlaps” in a replica space . For example $`Q_0(1;1^{})`$ quantify density - density and $`Q_2(1;1^{})`$ response - density overlaps respectively between two space-time points. The “entropy” (88) corresponds to the volume in the dynamical phase space when not only fields $`\rho _\alpha (1)`$ but also overlaps $`Q_\mathrm{a}(1;1^{})`$ are given. In a sense the “entropy” (88) is again the generalization of the entropy for the heteropolymer spanned in a replica space at the given set of “overlaps” . ### B The saddle point treatment Let us introduce the functional $`\stackrel{~}{F}\{\psi _\alpha (1);\mathrm{\Phi }_\mathrm{a}(1;1^{})\}`$ by the functional Fourier transformation $`\mathrm{exp}\{\stackrel{~}{W}\{\rho _\alpha (1);Q_\mathrm{a}(1;1^{})\}\}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{\alpha =0}{\overset{1}{}}}{\displaystyle \underset{a=0}{\overset{3}{}}}D\psi _\alpha (1)D\mathrm{\Phi }_\mathrm{a}(1;1^{})\mathrm{exp}\{\stackrel{~}{F}\{\psi _\alpha (1);\mathrm{\Phi }_\mathrm{a}(1;1^{})\}`$ (96) $`+`$ $`i{\displaystyle }d1\rho _\alpha (1)\psi _\alpha (1)+i{\displaystyle }d1d2Q_\mathrm{a}(1;2)\mathrm{\Phi }_a(1;2)\}.`$ (97) After substitution in eq.(85) and integration over $`\rho _\alpha `$ and $`Q_\mathrm{a}(1;2)`$ one gets $`Z_{\mathrm{av}}\{\chi _\alpha ,H_a\}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{\alpha =0}{\overset{1}{}}}{\displaystyle \underset{a=0}{\overset{3}{}}}D\psi _\alpha (1)D\mathrm{\Phi }_\mathrm{a}(1;1^{})\mathrm{exp}\{\stackrel{~}{F}\{\psi _\alpha (1);\mathrm{\Phi }_\mathrm{a}(1;1^{})\}`$ (98) $``$ $`{\displaystyle \frac{N}{2\xi _0}}{\displaystyle d1d2\psi _\alpha (1)\left[v^1\right]_{\alpha \beta }(1,2)\psi _\beta (2)}`$ (99) $``$ $`{\displaystyle \frac{N}{\chi _0^2}}{\displaystyle }d1d2d3d4\mathrm{\Phi }_\mathrm{a}(1;2)\left[\mathrm{\Gamma }^1\right]_{\mathrm{ab}}(1,2,3,4)\mathrm{\Phi }_\mathrm{b}(3;4)\},`$ (100) where $`\stackrel{~}{F}\{\psi _\alpha (1);\mathrm{\Phi }_\mathrm{a}(1;1^{})\}=\mathrm{log}{\displaystyle }{\displaystyle \underset{p=1}{\overset{N}{}}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]`$ (101) $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle 𝑑tr_\alpha ^{(p)}(t)\left[\psi _\alpha \left(𝐫^{(p)}(t)\right)+i\chi _\alpha \left(𝐫^{(p)}(t)\right)\right]}`$ (102) $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dtdt^{}p_a^{(p)}(t;t^{})[\mathrm{\Phi }_a[𝐫^{(p)}(t);𝐫^{(p)}(t^{})]+iH_a[𝐫^{(p)}(t);𝐫^{(p)}(t^{})]]\}.`$ (103) In order to ensure the extensivity of the whole effective action in eq.(100) we put the variance $`\chi ^2=\chi _0^2N`$ (so that the variance of the whole strength factor in eq.(4) scaled as $`N^{1/2}`$ akin to ref.) . This enable to represent GF in a similar to eq.(24) form $`Z_{\mathrm{av}}\left\{\chi _\alpha \right\}={\displaystyle \underset{\alpha =0}{\overset{1}{}}\underset{a=1}{\overset{3}{}}D\psi _\alpha (1)D\mathrm{\Phi }_a\mathrm{exp}\left\{N\stackrel{~}{A}[\psi _\alpha ,\mathrm{\Phi }_a;\chi _\alpha ,H_a]\right\}},`$ (104) where $`\stackrel{~}{A}[\psi _\alpha ,\mathrm{\Phi }_a;\chi _\alpha ,H_a]={\displaystyle \frac{1}{2\xi _0}}{\displaystyle 𝑑td1d2\psi _\alpha (1)\left[v^1\right]_{\alpha \beta }(1,2)\psi _\beta (2)}`$ (105) $`{\displaystyle \frac{1}{\chi _0^2}}{\displaystyle d1d2d3d4\mathrm{\Phi }_\mathrm{a}(1;2)\left[\mathrm{\Gamma }^1\right]_{\mathrm{ab}}(1,2,3,4)\mathrm{\Phi }_\mathrm{b}(3;4)}{\displaystyle \frac{1}{N}}\mathrm{log}{\displaystyle \underset{p=1}{\overset{N}{}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)}`$ (106) $`\times \mathrm{exp}\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dtr_\alpha ^{(p)}(t)[\psi _\alpha \left(𝐫^{(p)}(t)\right)+i\chi _\alpha \left(𝐫^{(p)}(t)\right)]`$ (107) $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dtdt^{}p_a^{(p)}(t;t^{})[\mathrm{\Phi }_a[𝐫^{(p)}(t);𝐫^{(p)}(t^{})]+iH_a[𝐫^{(p)}(t);𝐫^{(p)}(t^{})]]\}.`$ (108) The resulting SP - equation reads $`\overline{\psi _\alpha }(1)`$ $`=`$ $`{\displaystyle \frac{i\xi _0}{N}}{\displaystyle d2v_{\alpha \beta }(1,2)\rho _\beta (2)_{\mathrm{SP}}}`$ (109) $`\overline{\mathrm{\Phi }_a}(1)`$ $`=`$ $`{\displaystyle \frac{i\xi _0}{N}}{\displaystyle d3d4\mathrm{\Gamma }_{ab}(1,2,3,4)\rho _b(2)_{\mathrm{SP}}},`$ (110) where the average $`<\mathrm{}>_{SP}`$ is calculated with the GF $`Z_0_{\mathrm{av}}\{\chi _\alpha ,H_a\}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{p=1}{\overset{N}{}}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]`$ (111) $``$ $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle 𝑑tr_\alpha ^{(p)}(t)\left[\overline{\psi _\alpha }\left(𝐫^{(p)}(t)\right)+i\chi _\alpha \left(𝐫^{(p)}(t)\right)\right]}`$ (112) $``$ $`i{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }dtdt^{}p_a^{(p)}(t;t^{})[\overline{\mathrm{\Phi }_a}[𝐫^{(p)}(t);𝐫^{(p)}(t^{})]+iH_a[𝐫^{(p)}(t);𝐫^{(p)}(t^{})]]\}.`$ (113) Thereby we are left with the GF of a free system which experiences the external mean-fields $`\overline{\psi _\alpha }+i\chi _\alpha `$ and $`\overline{\mathrm{\Phi }_a}+iH_a`$. ### C The self-consistent Hartree approximation In order to calculate GF given by eq.(113) we will use the self-consistent Hartree approximation (SCHA). For this approximation we replace the real action by an appropriate Gaussian one in such a way that all terms which include more then two fields $`r_j^{(p)}(t)`$ or/and $`\widehat{r}_j^{(p)}(t)`$ are written in all possible ways as products of pairs of $`r_j^{(p)}(t)`$ or $`\widehat{r}_j^{(p)}(t)`$ coupled to self-consistent averages of the remaining fields. The analogy between SCHA and SP-approximation at $`N\mathrm{}`$ for the special case when the non-quadratic terms in the action are only the functions of the mean-squared displacement $`d^2(tt^{})=_{p=1}^N\left[𝐫^{(p)}(t)𝐫^{(p)}(t^{})\right]^2/N`$ has been proven in ref.. In our case the action in eq.(113) has a more general form. In the Appendix E we show that the SCHA and the next to the saddle point approximation (NSPA) merge and both become exact, if the GF with an arbitrary action can be treated by a steepest descent approach at $`N\mathrm{}`$. Let us make the Fourier transformation of the mean-fields $`\overline{\psi _\alpha }\left(𝐫^{(p)}(t)\right)`$ $`=`$ $`{\displaystyle \frac{d^3k}{(2\pi )^3}\overline{\psi _\alpha }\left(𝐤\right)\mathrm{exp}\left\{i\mathrm{𝐤𝐫}^{(p)}(t)\right\}}`$ (114) $`\overline{\mathrm{\Phi }_a}(𝐫^{(p)}(t);𝐫^{(p)}(t^{}))`$ $`=`$ $`{\displaystyle \frac{d^3k^1d^3k^2}{(2\pi )^6}\overline{\mathrm{\Phi }_a}(𝐤^\mathrm{𝟏},𝐤^\mathrm{𝟐})\mathrm{exp}\left\{i𝐤^\mathrm{𝟏}𝐫^{(p)}(t)+ı𝐤^\mathrm{𝟐}𝐫^{(p)}(t^{})\right\}}`$ (115) and insert it in eq.(113) . Then for eq.(113) we use the Hartree - type action (see eq.(D33). By doing so we put for simplicity the expectation value $`\xi _0=0`$ . It is easy to assure oneselves also that the “response \- response overlap” $`Q_1(1,1^{})=0`$ (similar to $`\widehat{\sigma }\widehat{\sigma }=0`$ in ref.). In the curse of the derivation we have used SP - equation (110) and defined the correlator (or the incoherent scattering function) $`C(𝐤^\mathrm{𝟏},t;𝐤^\mathrm{𝟐},t^{})={\displaystyle \frac{1}{N}}Q_0(𝐤^\mathrm{𝟏},t;𝐤^\mathrm{𝟐},t^{})`$ (116) as well as the response functions $`G(𝐤^\mathrm{𝟏},t;𝐤^\mathrm{𝟐},t^{})`$ $`=`$ $`{\displaystyle \frac{1}{N}}Q_3(𝐤^\mathrm{𝟏},t;𝐤^\mathrm{𝟐},t^{})\text{at}t^{}<t`$ (117) $`G(𝐤^\mathrm{𝟐},t;𝐤^\mathrm{𝟏},t^{})`$ $`=`$ $`{\displaystyle \frac{1}{N}}Q_2(𝐤^\mathrm{𝟏},t;𝐤^\mathrm{𝟐},t^{})\text{at}t^{}>t,`$ (118) where $`\mathrm{}`$ stands for the averaging with the Hartree \- type of action. After collection of all terms the final result (at $`\chi _\alpha =0\text{and}H_a=0`$) reads then $`Z_0_{\mathrm{av}}\{\overline{\psi _\alpha },\overline{\mathrm{\Phi }_a}\}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{p=1}{\overset{N}{}}}D𝐫^{(p)}(t)D\widehat{𝐫}^{(p)}(t)\mathrm{exp}\{{\displaystyle \underset{p=1}{\overset{N}{}}}A_0[𝐫^{(p)},\widehat{𝐫}^{(p)}]+{\displaystyle }dtdt^{}i\widehat{𝐫}^{(p)}(t)𝐫^{(p)}(t)\lambda (t,t^{})`$ (119) $``$ $`{\displaystyle }dtdt^{}i\widehat{𝐫}^{(p)}(t)𝐫^{(p)}(t^{})\lambda (t,t^{})+{\displaystyle }dtdt^{}i\widehat{𝐫}^{(p)}(t)i\widehat{𝐫}^{(p)}(t^{})\eta (t,t^{})\},`$ (120) where $`\lambda (t,t^{})={\displaystyle \frac{2}{3}}\chi _0^2{\displaystyle \frac{d^3k}{(2\pi )^3}k^2\left|v(k)\right|^2G(𝐤;t,t^{})C(𝐤;t,t^{})}`$ (121) and $`\eta (t,t^{})={\displaystyle \frac{1}{3}}\chi _0^2{\displaystyle \frac{d^3k}{(2\pi )^3}k^2\left|v(k)\right|^2\left[C(𝐤;t,t^{})\right]^2}.`$ (122) In eqs.(120) - (122) we have restricted ourselves to the homogeneous case $`C(𝐤,t;𝐤^{},t^{})`$ $`=`$ $`(2\pi )^3\delta (𝐤+𝐤^{})C(𝐤;t,t^{})`$ (123) $`G(𝐤,t;𝐤^{},t^{})`$ $`=`$ $`(2\pi )^3\delta (𝐤+𝐤^{})G(𝐤;t,t^{})`$ (124) for the correlation and response function. The equation of motion for the one particle correlator $`𝒫(t,t^{})={\displaystyle \frac{1}{3}}{\displaystyle \underset{j=1}{\overset{3}{}}}r_j^{(p)}(t)r_j^{(p)}(t^{})`$ (125) and the corresponding response function $`𝒢(t,t^{})={\displaystyle \frac{1}{3}}{\displaystyle \underset{j=1}{\overset{3}{}}}i\widehat{r}_j^{(p)}(t^{})r_j^{(p)}(t)`$ (126) (which actually does not depend from the particle index $`p`$) can be derived from eq.(120) by using the standard techniques . The resulting equations read $`\left[m_0{\displaystyle \frac{^2}{t^2}}+\gamma _0{\displaystyle \frac{}{t}}+{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau )\right]𝒫(t,t^{}){\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau )𝒫(t,t^{})`$ (127) $`+{\displaystyle \underset{\mathrm{}}{\overset{t^{}}{}}}𝑑\tau \eta (t,\tau )𝒢(t^{},\tau )=2T\gamma _0𝒢(t^{},t)`$ (128) and $`\left[m_0{\displaystyle \frac{^2}{t^2}}+\gamma _0{\displaystyle \frac{}{t}}+{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau )\right]𝒢(t,t^{}){\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau )𝒢(\tau ,t^{})`$ (129) $`=\delta (tt^{}).`$ (130) Eqs.(128) - (130) should be supplemented with the initial conditions $`\gamma _0𝒢(t+0^+,t)=1`$ and $`𝒢(t,t)=0`$. By making use this condition , equipartition $`(m_0/3)_{j=1}^3\dot{r}_j(t)\dot{r}_j(t)=T`$, causality $`𝒢(t,t^{})=0`$ at $`tt^{}`$ as well as the condition $`(1/3)_{j=1}^3\dot{r}_j(t)r_j(t)=0`$ one find from eq.(128) the following equation $`\left[{\displaystyle \frac{1}{2}}m_0{\displaystyle \frac{^2}{t^2}}+{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau )\right]𝒫(t,t){\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau )𝒫(\tau ,t^{})+{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \eta (t,\tau )𝒢(t,\tau )=2T.`$ (131) The set of eqs.(128) - (130) have the same structure as the Dyson eq.(49). After the matrices inversions and going to the time domain eqs.(49) (in time-translational invariant case) take the form $`\left[m_0{\displaystyle \frac{^2}{t^2}}+\gamma _0{\displaystyle \frac{}{t}}+\mu (0)\right]G_{01}(t,t^{})`$ $``$ $`{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \mathrm{\Sigma }_{10}(t\tau )G_{01}(\tau t^{})=\delta (tt^{})`$ (132) $`\left[m_0{\displaystyle \frac{^2}{t^2}}+\gamma _0{\displaystyle \frac{}{t}}+\mu (0)\right]G_{00}(t,t^{})`$ $``$ $`{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \mathrm{\Sigma }_{10}(t\tau )G_{00}(\tau t^{})`$ (133) $``$ $`{\displaystyle 𝑑\tau \mathrm{\Sigma }_{11}(t\tau )G_{10}(\tau t^{})}=2T\gamma _0G_{10}(tt^{}),`$ (134) where $`\mu (0)=\underset{0}{\overset{\mathrm{}}{}}𝑑t\lambda (t)`$ and RPA - Fourier spectrum $`S_{01}={\displaystyle \frac{1}{i\gamma _0\omega m_0\omega ^2+\mu (0)}}.`$ (135) Eqs.(128) - (130) are turned to the Dyson equations (133) - (134) provided that $`G_{00}(t)`$ $`=`$ $`𝒫(t),G_{01}(t)=𝒢(t),`$ (136) $`\mathrm{\Sigma }_{10}(t)`$ $`=`$ $`\lambda (t),\mathrm{\Sigma }_{11}(t)=\eta (t).`$ (137) We can show that the relation $`\beta {\displaystyle \frac{}{t}}\mathrm{\Sigma }_{11}(t)=\mathrm{\Sigma }_{10}(t)\mathrm{\Sigma }_{01}(t)`$ (138) holds, provided that the FDT is satisfied for $`G_{\alpha \beta }(t)`$. We then have in addition $`\beta {\displaystyle \frac{}{t}}G_{00}(t)=G_{01}(t)G_{10}(t).`$ (139) Bearing eqs.(137) in mind the eq.(138) takes in our case the form (at $`t>0`$) $`\beta {\displaystyle \frac{}{t}}\eta (t)=\lambda (t).`$ (140) The validity of the relationship (140) can be checked by replacing (121) and (122) in (140). The general eqs.(128) - (130) are equivalent, mutatis mutandis, to the corresponding equations for the $`p`$ \- spin system or a particle in the random potential at the large dimension . The most important features of these equations are the glassy dynamical behavior and the universal aging regime. At low temperatures the system tries to minimize the energy and each particle (with a number $`p`$) tends to surround itself with other particles which assure the strength parameter $`\mu _{\mathrm{pm}}<0`$. On the other hand the long - range interaction tries to support other pairs $`(ij)`$ corresponding to $`\mu _{ij}>0`$. As a result the system becomes “frustrated” and many local free energy minima appear. In the spirit of ref. when $`t,t^{}\mathrm{}`$ we have to discriminate between different cases: (i) the asymptotic regime when $`(tt^{})/t0`$ and (ii) the aging regime when $`(tt^{})/t^{}𝒪(1)`$. The aging regime is much more complicated because the time - translational invariance and FDT are violated. This regime has been extensively investigated both theoretically and by the computer simulation . In the following we restrict ourselves only to the asymptotic regime, for the sake of clearness and simplicity, and since the main features will be already visible. ### D The asymptotic regime This asymptotic regime is characterized by the large time scales, i.e., $`t,t^{}\mathrm{}`$ but keeping the difference $`\tau =tt^{}`$ finite. Under these circumstances we can define $`𝒫_{\mathrm{as}}(\tau )=\underset{t^{}\mathrm{}}{lim}𝒫(t^{}+\tau ,t^{})`$ (141) $`𝒢_{\mathrm{as}}(\tau )=\underset{t^{}\mathrm{}}{lim}𝒢(t^{}+\tau ,t^{})`$ (142) Then the equation for the displacement $`𝒟_{\mathrm{as}}=2[𝒫_{\mathrm{as}}(0)𝒫_{\mathrm{as}}(\tau )]`$, response function $`𝒢_{\mathrm{as}}(\tau )`$ and the static correlator $`𝒫_{\mathrm{as}}(0)`$ takes correspondingly the forms $`\left[m_0{\displaystyle \frac{^2}{\tau ^2}}+\gamma _0{\displaystyle \frac{}{\tau }}+M\right]𝒟_{\mathrm{as}}(\tau ){\displaystyle \underset{0}{\overset{\tau }{}}}𝑑\tau ^{}\lambda _{\mathrm{as}}(\tau \tau ^{})𝒟_{\mathrm{as}}(\tau ^{})`$ (143) $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑\tau ^{}\left[\lambda _{\mathrm{as}}(\tau +\tau ^{})\lambda _{\mathrm{as}}(\tau ^{})\right]𝒟_{\mathrm{as}}(\tau ^{})2{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑\tau ^{}\left[\eta _{\mathrm{as}}(\tau +\tau ^{})\eta _{\mathrm{as}}(\tau ^{})\right]𝒢_{\mathrm{as}}(\tau ^{})=2T`$ (144) $`\left[m_0{\displaystyle \frac{^2}{\tau ^2}}+\gamma _0{\displaystyle \frac{}{\tau }}+M\right]𝒢_{\mathrm{as}}(\tau ){\displaystyle \underset{0}{\overset{\tau }{}}}𝑑\tau ^{}\lambda _{\mathrm{as}}(\tau \tau ^{})𝒢_{\mathrm{as}}(\tau ^{})=0`$ (145) $`𝒫_{\mathrm{as}}(0)={\displaystyle \frac{1}{MM_{\mathrm{as}}}}\left[T{\displaystyle \frac{1}{2}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑\tau \lambda _{\mathrm{as}}𝒟_{\mathrm{as}}(\tau ){\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑\tau \eta _{\mathrm{as}}𝒢_{\mathrm{as}}(\tau )\right],`$ (146) where $`M=\underset{t\mathrm{}}{lim}{\displaystyle \underset{\mathrm{}}{\overset{t}{}}}𝑑\tau \lambda (t,\tau ).`$ (147) $`M_{\mathrm{as}}={\displaystyle \underset{0}{\overset{\mathrm{}}{}}}𝑑\tau \lambda _{\mathrm{as}}(\tau ).`$ (148) However, it is also convenient to define the “anomaly” $`\overline{M}=MM_{\mathrm{as}}`$ . The eqs.(144) - (146) has been analyzed first in the context of polymeric manifold in the random media and the random - phase sine - Gordon model . The peculiarity of our model is defined by its memory functions $`\lambda _{\mathrm{as}}(\tau )`$ and $`\eta _{\mathrm{as}}(\tau )`$. For example, let as give an explicit expression for $`\eta _{\mathrm{as}}(\tau )`$. The Gaussian form of the correlator, $`C(\tau )=\mathrm{exp}\left\{k^2𝒟_{\mathrm{as}}(\tau )/2\right\}`$, leads from eq.(122)to the result $`\eta _{\mathrm{as}}(\tau )={\displaystyle \frac{\chi _0^2\sqrt{\pi }}{6}}{\displaystyle \frac{1}{\sqrt{𝒟_{\mathrm{as}}(\tau )}}}.`$ (149) Usually it is assumed that at the high temperature FDT holds, i.e. $`\beta {\displaystyle \frac{}{\tau }}𝒟_{\mathrm{as}}(\tau )=2𝒢_{\mathrm{as}}(\tau )`$ (150) and $`\beta {\displaystyle \frac{}{\tau }}\eta _{\mathrm{as}}(\tau )=2\lambda _{\mathrm{as}}(\tau ).`$ (151) In this case eqs.(144) and (145) merge and take a simple form $`\left[m_0{\displaystyle \frac{^2}{\tau ^2}}+\gamma _0{\displaystyle \frac{}{\tau }}+M\right]𝒟_{\mathrm{as}}(\tau ){\displaystyle \underset{0}{\overset{\tau }{}}}𝑑\tau ^{}\lambda _{\mathrm{as}}(\tau \tau ^{})𝒟_{\mathrm{as}}(\tau ^{})=2T.`$ (152) It turns out that the solution which satisfies the FDT is only stable above a critical temperature $`T_\mathrm{c}`$. For the stability analysis it is convenient to represent eq.(152) in the form $`\left[m_0{\displaystyle \frac{^2}{\tau ^2}}+\gamma _0{\displaystyle \frac{}{\tau }}+\overline{M}+M_{\mathrm{as}}(\tau )\right]𝒟_{\mathrm{as}}(\tau ){\displaystyle \underset{0}{\overset{\tau }{}}}𝑑\tau ^{}\left[\eta _{\mathrm{as}}(\tau \tau ^{})\eta _{\mathrm{as}}(\tau )\right]{\displaystyle \frac{}{\tau ^{}}}𝒟_{\mathrm{as}}(\tau ^{})=2T,`$ (153) where $`M_{\mathrm{as}}(\tau )={\displaystyle \underset{\tau }{\overset{\mathrm{}}{}}}𝑑\tau ^{}\lambda _{\mathrm{as}}(\tau ^{}).`$ (154) For $`\tau \mathrm{}`$ the stability condition which comes out of eq.(153) reads $`\left[\overline{M}+M_{\mathrm{as}}(\tau )\right]𝒟_{\mathrm{as}}(\tau )2T.`$ (155) Then the stationary value of the displacement $`𝒟_{\mathrm{as}}(\tau \mathrm{})=q_0`$ reads $`q_0={\displaystyle \frac{2T}{\overline{M}}}.`$ (156) By taking into account eqs.(150) and (151) the stability condition becomes $`D(q,T)0`$ (157) for $`0qq_0`$, where $`D(q,T)\left[\left({\displaystyle \frac{\chi _0}{T}}\right)^2{\displaystyle \frac{\sqrt{\pi }}{12\sqrt{q_0}}}{\displaystyle \frac{1}{q_0}}\right]q\left({\displaystyle \frac{\chi _0}{T}}\right)^2{\displaystyle \frac{\sqrt{\pi }}{12}}\sqrt{q}+1.`$ (158) The critical values $`q_\mathrm{c}`$ and $`T_\mathrm{c}`$ at which the condition (157) first becomes violated is defined by equations $`D(q_\mathrm{c},T_\mathrm{c})`$ $`=`$ $`0`$ (159) $`D^{}(q_\mathrm{c},T_\mathrm{c})`$ $`=`$ $`0.`$ (160) Consequently, eqs. (160) have the simple solution $`\left({\displaystyle \frac{T_\mathrm{c}}{\chi _0}}\right)^2={\displaystyle \frac{\sqrt{\pi q_0}}{24}}\text{and}q_\mathrm{c}=q_0.`$ (161) The Fig.3 shows the behavior of $`D(q,T)`$ in the vicinity of the critical point. It can be seen that the minimum, $`q_mq_0`$, at which $`D(q,T)0`$ appears continuously, i.e. the instability of the FDT solution shows up as a $`2^{\mathrm{nd}}`$ order phase transition. This is analogous to the dynamics of polymeric manifolds in a medium with the long range correlation in disorder . In particular, if “anomaly” $`\overline{M}0`$ then $`q_0\mathrm{}`$ and $`T_\mathrm{c}\mathrm{}`$, so in this case the FDT solution is unstable for any finite temperature. Let us consider the dynamics at the temperature slightly above the critical point: $`T=T_\mathrm{c}(1+\epsilon )`$, where $`0<\epsilon 1`$. For large $`\tau `$ the decomposition $`𝒟_{\mathrm{as}}(\tau )=q_0f(\tau ),`$ (162) is possible, where $`f(\tau )q_0`$. The substitution of this decomposition into eq.(153) and the expansion up to the second order with respect to $`f(\tau )`$ yields $`\epsilon q_0f(\tau )+{\displaystyle \frac{1}{8}}\left[f(\tau )\right]^2+{\displaystyle \frac{1}{2}}{\displaystyle \underset{0}{\overset{\tau }{}}}𝑑\tau ^{}\left[f(\tau \tau ^{})f(\tau )\right]{\displaystyle \frac{}{\tau ^{}}}f(\tau ^{})=0.`$ (163) Following ref. let us make the Laplace transformation $`[f(\tau )]\stackrel{~}{f}(z)`$ and introduce the scaling functions, $`\widehat{\varphi }(\widehat{z})`$ or $`\widehat{\varphi }(\widehat{\tau })`$ , in a such way $`\stackrel{~}{f}(z)={\displaystyle \frac{c_\epsilon }{\omega _\epsilon }}\widehat{\varphi }(\widehat{z})\text{or}f(\tau )=c_\epsilon \widehat{\varphi }(\widehat{\tau }),`$ (164) where $`\widehat{z}=z/\omega _\epsilon `$ and $`\widehat{\tau }=\tau \omega _\epsilon `$. If $`c_\epsilon =\epsilon `$ and $`\omega _\epsilon =\omega _0\epsilon ^{1/a}`$ then one can write eq.(163) in the form $`q_0\widehat{\varphi }(\widehat{z}){\displaystyle \frac{3}{8}}\left\{\widehat{\varphi }^2(\widehat{\tau })\right\}(\widehat{z})+\widehat{z}\widehat{\varphi }^2(\widehat{z})=0`$ (165) (see eq.(2.68b) from ref.). In the critical regime $`\widehat{z}1`$ (or $`\widehat{\tau }1`$) the solution of eq.(165) has a form $`\widehat{\varphi }(\widehat{\tau })\widehat{\tau }^a`$. In this limit the first term in eq.(165) is dropped out and the exponent is defined by the equation $`{\displaystyle \frac{\mathrm{\Gamma }^2(1a)}{\mathrm{\Gamma }(12a)}}={\displaystyle \frac{3}{4}}.`$ (166) The solution of eq. (166) gives $`a=0.30465`$. In the opposite limit $`\widehat{z}1`$ (or $`\widehat{\tau }1`$) the last term in eq.(165) can be neglected. In this case the solution has the form $`\widehat{\varphi }(\widehat{\tau })A_\epsilon \widehat{\tau }^a\mathrm{exp}\{\lambda \widehat{\tau }\}`$, where $`A_\epsilon =8\epsilon q_0\mathrm{\Gamma }(1a)2^{(12a)}/3\mathrm{\Gamma }(12a)\lambda ^2`$. As a result the overall scaling reads $`𝒟_{\mathrm{as}}(\tau )=\{{\displaystyle \genfrac{}{}{0pt}{}{q_0\frac{c_\epsilon }{(\omega _\epsilon \tau )^a},\text{at}\omega _\epsilon \tau 1}{q_0\frac{A_\epsilon }{(\omega _\epsilon \tau )^a}\mathrm{exp}\{\lambda (\omega _\epsilon \tau )\},\text{at}\omega _\epsilon \tau 1}},`$ (167) where $`\lambda `$ is some constant. At $`T<T_\mathrm{c}`$ the FDT is violated for the large time separation $`\tau `$ and the aging regime is arising. It should be mentioned that the asymptotic regime cannot be decoupled from the aging one . In actual fact, the “anomaly” $`\overline{M}`$ in the asymptotic eqs. (144) - (146) strictly speaking can be calculated only from the aging regime. Because of the distinct aim in this paper we are not going to discuss the aging regime here expecting to return to it in a later publication. ## IV Conclusion In the present paper we have considered the dynamics of two models with the long range repulsive interaction. The interaction potential was designed in a way to enable the saddle point treatment, as well as a fluctuation expansion. For the pure model we have derived eq.(61) for the full correlation matrix $`G_{\alpha \beta }(𝐤,\omega )`$ in the one-loop approximation, which has an the explicit solution (see eqs.(65) - (67) ). This solution has a “boring behavior” at $`\omega 0`$ which manifests the absence of the glass dynamics. The physical background of this stems from the fact that the potential is much too soft and the “cage effect” is completely missing. This conclusion is in accordance with the interacting particles statistical thermodynamics analysis, which was given in ref.. It was shown there that for the infinite range interaction potential, which allows a well defined saddle point treatment, the glassy phase is simply suppressed. On the other hand, the same model but with a randomly distributed strength of interaction (the “random - bond model”) leads to the continuous glass transition. This type of transition is also the case for the polymeric manifolds in the disordered medium with long range correlation as well as for the $`p`$ \- spin interaction spin-glass model at the large external field . It would be also interesting to investigate the more realistic “random sequence model” when each particle carry a random “charge”. Qualitatively the same glassy behavior have been found in the pure spin models with the deterministic but very rapidly oscillating coupling between variables . It was assumed that the effective quenched disorder is “self - induced”. This means that because of the slow dynamics some degrees of freedom freeze and play the role of the effectively quenched disorder. As a conclusion, the glass transition in the pure systems of the interacting particles, where the disorder is actually “self - induced”, goes beyond the mean-field level . This appears to hard to implement in the present context, because it implies the consideration of the short range interaction potential as well as activated processes. ## V Acknowledgments The authors have benefited from discussions with M. Fuchs, A. Latz, G. Migliorini and greatly indebted to the Deutsche Forschungsgemeinschaft (DFG), the Sonderforschungsbereich SFB 262 for financial support of the work. ## A Calculation of the 4-point RPA - correlation matrix In the full analogy with eq.(32) the expression for $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ reads $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)=\underset{\overline{\psi _\alpha }+i\chi _\alpha .0}{lim}\left[{\displaystyle \frac{\delta }{N^2\delta \chi _\beta (2)\delta \chi _\gamma (3)\delta \chi _\delta (4)}}\rho _\alpha (1)_{\mathrm{SP}}\right]`$ (A1) The expansion of the $`\rho _\alpha (1)_{\mathrm{SP}}`$ up to the $`3^\mathrm{d}`$ order with respect to the mean field $`\overline{\psi _\alpha }+i\chi _\alpha `$ can be easily obtained from eq.(31) $`\rho _\alpha (1)_{\mathrm{SP}}`$ $`=`$ $`\rho _\alpha (1)_0+{\displaystyle d2\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)_0\left[\chi _\beta (2)i\overline{\psi _\beta }(2)\right]}`$ (A2) $`+`$ $`{\displaystyle \frac{1}{2!}}{\displaystyle d2d3\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)\mathrm{\Delta }\rho _\gamma (3)_0\left[\chi _\beta (2)i\overline{\psi _\beta }(2)\right]\left[\chi _\gamma (3)i\overline{\psi _\gamma }(3)\right]}`$ (A3) $`+`$ $`{\displaystyle \frac{1}{3!}}{\displaystyle d2d3d4\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)\mathrm{\Delta }\rho _\gamma (3)\mathrm{\Delta }\rho _\delta (4)_0}`$ (A4) $`\times `$ $`\left[\chi _\beta (2)i\overline{\psi _\beta }(2)\right]\left[\chi _\gamma (3)i\overline{\psi _\gamma }(3)\right]\left[\chi _\delta (4)i\overline{\psi _\delta }(4)\right].`$ (A5) By using the SP-equation (29) and after threefold differentiation with respect to $`\chi _\alpha (1)`$ (see eq.(A1)) we find $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ $`=`$ $`F_{\overline{\alpha }\overline{\beta }\overline{\gamma }\overline{\delta }}^{(4)}(\overline{1},\overline{2},\overline{3},\overline{4})`$ (A6) $`\times `$ $`\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{F}^{(2)}\right]^1\right\}_{\overline{\alpha }\alpha }(\overline{1},1)\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{F}^{(2)}\right]^1\right\}_{\overline{\beta }\beta }(\overline{2},2)`$ (A7) $`\times `$ $`\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{F}^{(2)}\right]^1\right\}_{\overline{\gamma }\gamma }(\overline{3},3)\left\{\left[\widehat{1}+\mu \widehat{v}\widehat{F}^{(2)}\right]^1\right\}_{\overline{\delta }\delta }(\overline{4},4),`$ (A8) where the 4-point free system correlation matrix $`F_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)={\displaystyle \frac{1}{N^2}}\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)\mathrm{\Delta }\rho _\gamma (3)\mathrm{\Delta }\rho _\delta (4)_0.`$ (A9) In eq.(A8) we imply the summation (integration) over the barred indices (barred space - time variables). When deriving eq.(A8) we have also kept in mind that $`\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)N`$ and $`\mathrm{\Delta }\rho _\alpha (1)\mathrm{\Delta }\rho _\beta (2)\mathrm{\Delta }\rho _\gamma (3)\mathrm{\Delta }\rho _\delta (4)_0N^2`$ , etc. The fact that the matrix $`F_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ is symmetrical with respect to simultaneous permutations of Greek indices and space-time arguments as well as eq.(33) have been used. It is easy to show that $`F_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ is factorized $`F_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ $`=`$ $`F_{\alpha \beta }^{(2)}(1,2)F_{\gamma \delta }^{(2)}(3,4)+F_{\alpha \gamma }^{(2)}(1,3)F_{\beta \delta }^{(2)}(2,3)+F_{\alpha \delta }^{(2)}(1,4)F_{\beta \gamma }^{(2)}(2,3).`$ (A10) On the other side it is instructive to check that even in this case $`S_{\alpha \beta \gamma \delta }^{(4)}(1,2,3,4)`$ can not be factorized. ## B calculation of the vertex-matrix $`K_{\alpha \beta }(1,2)`$ The substitution of the eg.(A8) into eq.(59) after the straightforward algebra yields $`K_{\alpha \beta }(1,2)`$ $`=`$ $`\left\{\left[\left(2\mu \right)^1\widehat{v}^1+\widehat{F}\right]^1\left[\mu ^1\widehat{v}^1+\widehat{F}\right]^1\right\}_{\overline{\beta }\overline{\alpha }}(\overline{2},\overline{1})`$ (B1) $`\times `$ $`\left\{F_{\overline{\alpha }\overline{\beta }}(\overline{1},\overline{2})F_{\overline{\gamma }\overline{\delta }}(\overline{3},\overline{4})+F_{\overline{\alpha }\overline{\gamma }}(\overline{1},\overline{3})F_{\overline{\beta }\overline{\delta }}(\overline{2},\overline{4})+F_{\overline{\alpha }\overline{\delta }}(\overline{1},\overline{4})F_{\overline{\beta }\overline{\gamma }}(\overline{2},\overline{3})\right\}`$ (B2) $`\times `$ $`\left\{\left[\overline{1}+2\mu \overline{v}\overline{F}\right]^1\right\}_{\overline{\gamma }\alpha }(\overline{3},1)\left\{\left[\overline{1}+2\mu \widehat{v}\widehat{F}\right]^1\right\}_{\overline{\delta }\beta }(\overline{4},2),`$ (B3) where as before for the repeated barred indices (variables) the summation (integration) is implied. For the time - space - translational invariant case the respective Fourier transformation leads to the result: $`K_{\alpha \beta }(𝐤,\omega )`$ $`=`$ $`\{IF_{\overline{\gamma }\overline{\delta }}(𝐤,\omega )`$ (B4) $`+`$ $`\left\{\left[\left(2\mu \right)^1\widehat{v}^1+\widehat{F}\right]^1\left[\mu ^1\widehat{v}^1+\widehat{F}\right]^1\right\}_{\overline{\beta }\overline{\alpha }}(𝐤,\omega )F_{\overline{\alpha }\overline{\gamma }}(𝐤,\omega )F_{\overline{\beta }\overline{\delta }}(𝐤,\omega )`$ (B5) $`+`$ $`\{[\left(2\mu \right)^1\widehat{v}^1+\widehat{F}]^1[\mu ^1\widehat{v}^1+\widehat{F}]^1\}_{\overline{\beta }\overline{\alpha }}(𝐤,\omega )F_{\overline{\alpha }\overline{\delta }}(𝐤,\omega )F_{\overline{\beta }\overline{\gamma }}(𝐤,\omega )\}`$ (B6) $`\times `$ $`\left\{\left[\overline{1}+2\mu \widehat{v}\widehat{F}\right]^1\right\}_{\overline{\gamma }\alpha }(𝐤,\omega )\left\{\left[\overline{1}+2\mu \overline{v}\overline{F}\right]^1\right\}_{\overline{\delta }\beta }(𝐤,\omega ),`$ (B7) where the trace $`I={\displaystyle \frac{d^3qd\omega }{(2\pi )^4}\left\{\left[\widehat{1}+\left(2\mu \right)^1\widehat{F}^1\widehat{v}^1\right]^1\left[\widehat{1}+\mu ^1\widehat{F}^1\widehat{v}^1\right]^1\right\}_{\overline{\alpha }\overline{\alpha }}(𝐪,\omega )}.`$ (B8) With the correlation matrix $`\widehat{F}`$ given by eqs.(64), by doing integration over $`\omega `$ one can check that the trace $`I=0`$. This gives finally $`K_{\alpha \beta }(𝐤,\omega )=L_{\alpha \beta }(𝐤,\omega )+L_{\alpha \beta }(𝐤,\omega ),`$ (B9) where $`L_{\alpha \beta }(𝐤,\omega )`$ $`=`$ $`\left\{\left[\overline{1}+2\mu \widehat{v}\widehat{F}\right]^1\right\}_{\alpha \overline{\gamma }}(𝐤,\omega )F_{\overline{\gamma }\overline{\beta }}(𝐤,\omega )`$ (B10) $`\times `$ $`\left\{\left[\widehat{1}+\left(2\mu \right)^1\widehat{F}^1\widehat{v}^1\right]^1\left[\widehat{1}+\mu ^1\widehat{F}^1\widehat{v}^1\right]^1\right\}_{\overline{\beta }\overline{\delta }}(𝐪,\omega )`$ (B11) $`\times `$ $`\left\{\left[\overline{1}+2\mu \widehat{v}\widehat{F}\right]^1\right\}_{\alpha \overline{\gamma }}(𝐤,\omega ).`$ (B12) ## C The MCT for the generalized Kac potential In this case the direct correlation function $`c(𝐫)=\beta V(𝐫)`$ and its Fourier transformation takes the scaling form $`c(𝐤)=\beta f\left({\displaystyle \frac{𝐤}{\kappa }}\right).`$ (C1) Let us insert this expression to the MCT - memory kernel (see eq.(3.32) in ). It is reasonable then to rescale the integration variables in the memory kernel , $`𝐤\kappa 𝐤,𝐩\kappa 𝐩`$ as well as to put for the external wave vector $`𝐪=\kappa 𝐪_0`$ , where $`𝐪_0`$ is some reference wave vector. The last scaling means that in the MF - limit an experiment probes a very small wave vector : $`𝐪0`$. The resulting scaling of the memory kernel , $`m(𝐪,t)`$, reads $`m(\kappa 𝐪_0,t)`$ $`=`$ $`\kappa ^d\stackrel{~}{S}(𝐪_0){\displaystyle \frac{\rho _0}{2}}{\displaystyle \frac{d𝐤d𝐩}{(2\pi )^{2d}}\delta ^{(d)}(𝐤+𝐩𝐪_0)}`$ (C2) $`\times `$ $`{\displaystyle \frac{\left\{e^L(𝐪)\beta \left[𝐤f(𝐤)+𝐩f(𝐩)\right]\right\}^2}{q_0^2}}\stackrel{~}{S}(𝐤,t)\stackrel{~}{S}(𝐩,t),`$ (C3) where we have took into account the scaling form of the correlator: $`S(𝐤,t)=\stackrel{~}{S}(𝐤/\kappa ;t)`$. Thus we finally arrive at the scaling $`m(\kappa 𝐪_0,t)\kappa ^d0`$ and the glass transition dies out. ## D The analogy between SCHA and NSPA Let us prove that SCHA becomes exact for GF given by (113) in the limit $`N\mathrm{}`$. We will consider even a more general GF $`Z\{\chi _\alpha \}`$ $`=`$ $`{\displaystyle }{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle \underset{\alpha =0,1}{}}Dx_\alpha ^{(p)}(1)\mathrm{exp}\{{\displaystyle \frac{1}{2}}{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }d1d2x_\alpha ^{(p)}(1)A_{\alpha \beta }(1,2)x_\beta ^{(p)}(2)`$ (D1) $`+`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}W\left[x_\alpha ^{(p)}\right]+{\displaystyle \underset{p=1}{\overset{N}{}}}{\displaystyle }d1x_\alpha ^{(p)}(1)\chi _\alpha (1)\},`$ (D2) where we have used the shorthand notations $`x_\alpha ^{(p)}(1)=\left(\begin{array}{c}r_j(t)\\ i\widehat{r}_j(t)\end{array}\right),`$ (D5) and “1” embraces Cartesian indices as well as time variable: $`1\{i,j,k;t\}`$. In eq.(D2) $`W\left[x_\alpha ^{(p)}\right]`$ is an arbitrary functional of $`x_\alpha ^{(p)}`$. Instead of the exact action functional in eq.(D2) we consider now the trial one which has a Gaussian form $`S\left[x_\alpha ^{(p)}(1)\right]`$ $`=`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}\{{\displaystyle \frac{1}{2}}{\displaystyle }d1d2x_\alpha ^{(p)}(1)A_{\alpha \beta }(1,2)x_\beta ^{(p)}(2)`$ (D6) $``$ $`{\displaystyle }d1d2x_\alpha ^{(p)}(1)\mathrm{\Gamma }_{\alpha \beta }(1,2)x_\beta ^{(p)}(2){\displaystyle }d1L_\alpha (1)x_\alpha ^{(p)}(1)\}.`$ (D7) Let us look for the “best” coefficients $`\mathrm{\Gamma }_{\alpha \beta }(1,2)`$ and $`L_\alpha (1)`$ in a sense that the exact “free energy” $`F\left[\chi _\alpha \right]=\mathrm{log}Z\left\{\chi _\alpha \right\}`$ tends to the trial one $`F_0\left[\chi _\alpha \right]=\mathrm{log}Dx_\alpha ^{(p)}\mathrm{exp}\{S[x_\alpha ^{(p)}]\}`$, i. e. $`F\left[\chi _\alpha \right]F_0\left[\chi _\alpha \right]`$ (D8) and both becomes exact at $`N\mathrm{}`$. We can show that the property (D7) is satisfied by $`\mathrm{\Gamma }_{\alpha \beta }`$ and $`L_\alpha `$ which are obtained by extremization of the functional $`\mathrm{\Phi }\{\mathrm{\Gamma }_{\alpha \beta },L_\alpha \}`$ $`=`$ $`\mathrm{log}{\displaystyle \underset{p=1}{\overset{N}{}}\underset{\alpha =0,1}{}Dx_\alpha ^{(p)}(1)\mathrm{exp}\left\{S\left[x_\alpha ^{(p)}\right]\right\}}`$ (D9) $`+`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}\{{\displaystyle }d1d2\mathrm{\Gamma }_{\alpha \beta }(1,2)x_\alpha ^{(p)}(1)x_\beta ^{(p)}(2)_\mathrm{s}`$ (D10) $`+`$ $`{\displaystyle }d1[L_\alpha (1)\chi _\alpha ]x_\alpha ^{(p)}(1)_\mathrm{s}W\left[x_\alpha ^{(p)}(1)\right]_\mathrm{s}\},`$ (D11) where we use the notations $`\mathrm{}_\mathrm{s}={\displaystyle \frac{_{p=1}^N_{\alpha =0,1}Dx_\alpha ^{(p)}(1)\mathrm{}\mathrm{exp}\left\{S\left[x_\alpha ^{(p)}\right]\right\}}{_{p=1}^N_{\alpha =0,1}Dx_\alpha ^{(p)}(1)\mathrm{exp}\left\{S\left[x_\alpha ^{(p)}\right]\right\}}}.`$ (D12) The extremization conditions reads $`{\displaystyle \frac{\delta }{\delta \mathrm{\Gamma }_{\alpha \gamma }(1,2)}}\mathrm{\Phi }`$ $`=`$ $`0`$ (D13) $`{\displaystyle \frac{\delta }{\delta L_\alpha (1)}}\mathrm{\Phi }`$ $`=`$ $`0.`$ (D14) The variations in eqs.(D12) can be done directly. During the calculation the generalized Wick’s theorem should be also taken into account. Namely, because the averaging (D11) is simply the Gaussian integral the Wick’s theorem yields $`x_\alpha ^{(p)}(1)W\left[x_\alpha ^{(p)}\right]_\mathrm{s}`$ $`=`$ $`x_\alpha ^{(p)}(1)_\mathrm{s}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}`$ (D15) $`+`$ $`{\displaystyle d2\mathrm{\Delta }x_\alpha ^{(p)}(1)\mathrm{\Delta }x_\beta ^{(p)}(2)_\mathrm{s}\frac{\delta }{\delta x_\beta ^{(p)}(2)}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}},`$ (D16) where $`\mathrm{\Delta }x_\alpha ^{(p)}(1)x_\alpha ^{(p)}(1)x_\alpha ^{(p)}(1)_\mathrm{s}`$. After the straightforward calculation we find $`\mathrm{\Gamma }_{\alpha \gamma }(1,2)={\displaystyle \frac{1}{2}}{\displaystyle \frac{\delta ^2}{\delta x_\alpha ^{(p)}(1)\delta x_\beta ^{(p)}(2)}}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}`$ (D17) and $`L_\alpha (1)={\displaystyle d2\left[A_{\alpha \beta }(1,2)\mathrm{\Gamma }_{\alpha \beta }(1,2)\right]x_\alpha ^{(p)}(2)_\mathrm{s}}.`$ (D18) Then equations for the two moments take the form $`x_\alpha ^{(p)}(1)_\mathrm{s}={\displaystyle d2\left[A^1\right]_{\alpha \beta ]}(1,2)\left[\frac{\delta }{\delta x_\beta ^{(p)}(2)}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}+\chi _\beta (2)\right]}`$ (D19) and $`\mathrm{\Delta }x_\alpha ^{(p)}(1)\mathrm{\Delta }x_\beta ^{(p)}(2)_\mathrm{s}=\left\{\left[\widehat{A}2\widehat{\mathrm{\Gamma }}\right]^1\right\}_{\alpha \beta }(1,2),`$ (D20) where $`\widehat{A}`$ and $`\widehat{\mathrm{\Gamma }}`$ stands for the corresponding $`2\times 2`$ \- matrices. On the other side the saddle point (SP) treatment of eq.(D2) at $`N\mathrm{}`$ yields $`{\displaystyle d2A_{\alpha \beta }(1,2)\overline{x}_\beta ^{(p)}(2)}+{\displaystyle \frac{\delta W}{\delta x_\alpha ^{(p)}(1)}}|_{x_\alpha =\overline{x}_\alpha }+\chi _\alpha (1)=0`$ (D21) and $`\mathrm{\Delta }x_\alpha ^{(p)}(1)\mathrm{\Delta }x_\beta ^{(p)}(2)_{\mathrm{SP}}=\left\{\left[\widehat{A}2\widehat{B}\right]^1\right\}_{\alpha \beta }(1,2),`$ (D22) where $`B_{\alpha \beta }(1,2)={\displaystyle \frac{1}{2}}{\displaystyle \frac{\delta ^2W}{\delta x_\alpha ^{(p)}(1)\delta x_\beta ^{(p)}(2)}}|_{x_\alpha =\overline{x}_\alpha }`$ (D23) and $`\overline{x}_\alpha ^{(p)}(1)`$ stands for the field in SP. In order to show the analogy between eqs.(D19)-(D20) and eqs.(D21)-(D22) let us make the functional Fourier transformation $`\mathrm{exp}\left\{K\left[y_\alpha ^{(p)}(1)\right]\right\}={\displaystyle Dx_\alpha ^{(p)}(1)\mathrm{exp}\left\{W\left[x_\alpha ^{(p)}(1)\right]id1x_\alpha ^{(p)}(1)y_\alpha ^{(p)}(1)\right\}}`$ (D24) and its inversion $`\mathrm{exp}\left\{W\left[x_\alpha ^{(p)}(1)\right]\right\}={\displaystyle Dy_\alpha ^{(p)}(1)\mathrm{exp}\left\{K\left[y_\alpha ^{(p)}(1)\right]+id1x_\alpha ^{(p)}(1)y_\alpha ^{(p)}(1)\right\}}.`$ (D25) Then eqs.(D21) - (D22) can be written as $`\overline{x}_\alpha ^{(p)}(1)={\displaystyle d2\left[A^1\right]_{\alpha \beta }(1,2)\left[iy_\beta ^{(p)}(2)_{\mathrm{SP}}+\chi _\beta (2)\right]}`$ (D26) and $`\mathrm{\Delta }x_\alpha ^{(p)}(1)\mathrm{\Delta }x_\beta ^{(p)}(2)_{\mathrm{SP}}=\left\{\left[\widehat{A}+\mathrm{\Delta }y\mathrm{\Delta }y_{\mathrm{SP}}\right]^1\right\}_{\alpha \beta }(1,2),`$ (D27) where the correlation matrix $`\mathrm{\Delta }y\mathrm{\Delta }y_{\mathrm{SP}}=\mathrm{\Delta }y_\alpha ^{(p)}(1)\mathrm{\Delta }y_\beta ^{(p)}(2)_{\mathrm{SP}}`$ (D28) and $`\mathrm{}_{\mathrm{SP}}={\displaystyle \frac{Dy_\alpha ^{(p)}\mathrm{}\mathrm{exp}\left\{K\left[y_\alpha ^{(p)}\right]+id1\overline{x}_\alpha ^{(p)}(1)y_\alpha ^{(p)}(1)\right\}}{Dy_\alpha ^{(p)}\mathrm{exp}\left\{K\left[y_\alpha ^{(p)}\right]+id1\overline{x}_\alpha ^{(p)}(1)y_\alpha ^{(p)}(1)\right\}}}.`$ (D29) At $`N\mathrm{}`$ by making use eq.(D25) one can immediately see that $`{\displaystyle \frac{\delta }{\delta x_\alpha ^{(p)}(1)}}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}iy_\alpha ^{(p)}(1)_{\mathrm{SP}}`$ (D30) and $`{\displaystyle \frac{\delta ^2}{\delta x_\alpha ^{(p)}(1)\delta x_\beta ^{(p)}(2)}}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}\mathrm{\Delta }y_\alpha ^{(p)}(1)\mathrm{\Delta }y_\beta ^{(p)}(2)_{\mathrm{SP}},`$ (D31) and SCHA exactly corresponds to NSPA . For the case which was treated in Sec.III C $`x_\alpha ^{(p)}(1)_\mathrm{s}=0`$ and the Hartree - type action (D7) cast the form $`S\left[x_\alpha ^{(p)}(1)\right]`$ $`=`$ $`{\displaystyle \underset{p=1}{\overset{N}{}}}\{{\displaystyle \frac{1}{2}}{\displaystyle }d1d2x_\alpha ^{(p)}(1)A_{\alpha \beta }(1,2)x_\beta ^{(p)}(2)`$ (D32) $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle }d1d2{\displaystyle \frac{\delta ^2}{\delta x_\alpha ^{(p)}(1)\delta x_\beta ^{(p)}(2)}}W\left[x_\alpha ^{(p)}\right]_\mathrm{s}x_\alpha ^{(p)}(1)x_\beta ^{(p)}(2)\}.`$ (D33)
warning/0003/hep-ph0003059.html
ar5iv
text
# Critical couplings for chiral symmetry breaking via instantons ## I introduction Nonperturbative gauge dynamics present many challenges that are yet to be fully understood. One of these is the phenomenon of chiral symmetry breaking. A theory with an $`SU(N_c)`$ gauge symmetry and $`N_f`$ fermions in the fundamental representation has a global chiral symmetry in the massless limit. Provided that $`N_f<5.5N_c`$ this theory is asymptotically free. Close to this threshold the coupling is bounded above due to an infrared fixed point in the beta function. As $`N_f`$ becomes smaller the fixed point value increases. Below a critical value for $`N_f`$ the coupling can grow strong enough to break the chiral symmetry down to the vector symmetry: $$SU(N_f)_R\times SU(N_f)_L\times U(1)_VSU(N_f)_V\times U(1)_V.$$ (1) As a result the fermions acquire a dynamical mass, which acts as an order parameter for the spontaneous breaking of the chiral symmetry. This much is well understood, but the respective parts played by gauge exchanges and instantons in this process still needs clarification. It has been known for some time that gauge exchanges can generate a dynamical mass. To leading order (ladder approximation) in Landau gauge the critical coupling associated with gauge exchange dynamics is $$\alpha _c^{ge}=\frac{2\pi N_c}{3(N_c^21)}.$$ (2) Assuming a two-loop beta function, one can show that the critical number of flavors, below which dynamical chiral symmetry breaking occurs, is $$N_f^c=4N_c\frac{6N_c}{25N_c^215}4N_c.$$ (3) It is known that instanton dynamics can also break the chiral symmetry . Attempts to investigate this possibility have been hampered by the presence of an infrared divergence in the integral over the instanton size. In their recent investigation of the critical number of flavors for chiral symmetry breaking via instantons, Appelquist and Selipsky avoided the problem of infrared divergences by introducing the following two modifications of the Carlitz and Creamer gap equation : * They introduced a mass dependence for the fermion determinant that is not only valid for the small mass limit but also for the large mass limit . This function simply makes a sharp transition from the small mass behavior to the large mass behavior. The effect is to introduce a fermion mass scale below which the fermions are integrated out. * In Reference all couplings in the instanton expression are allowed to run according to the two-loop beta function. For those values of $`N_f`$ where the beta function has a fixed point the coupling remains fairly constant until it reaches the fermion mass scale. Fermions are integrated out below the fermion mass scale. This causes the beta function to revert back to normal asymptotic free running, which gives a suppression of the integrand in the infrared region because of the way it depends on the coupling. The result of their investigation was that the critical number of flavors for instantons is about $`4.77N_c`$. From this they concluded that the instanton contribution to chiral symmetry breaking is comparable to that of gauge exchanges. In this paper we reinvestigate chiral symmetry breaking through instantons, but our investigation differs from that of Reference in two essential aspects: * The recently proposed instanton effective action formalism is used to determine the conditions under which the symmetric vacuum becomes unstable. This formalism reproduces the Carlitz and Creamer gap equation in the small mass limit for $`SU(2)`$. However, we are not interested in the shape of the mass function, but rather in the critical couplings and the critical number of flavors. * We concentrate on continuous phase transitions, such as found for chiral symmetry breaking through gauge exchanges. The effective action formalism makes it clear that the instanton contribution responsible for this type of transition is not the one–instanton amplitude of Figure 1a but rather the instanton–anti-instanton amplitude with two mass insertions, shown in Figure 1b. The latter point needs clarification. The effective action can be interpreted as an effective potential. This makes it easier to determine when the stability of the symmetric vacuum is affected by the dynamics. For small couplings the global minimum of the effective potential is located at the origin, where the order parameter is zero. When the coupling is increased the global minimum remains at the origin until the coupling crosses the critical value where the phase transition occurs. At this point the global minimum can either jump discontinuously to a nonzero value of the order parameter, indicating a first order phase transition, or it can move continuously (though non-analytically) to a nonzero value of the order parameter, indicating a continuous phase transition. For a first order phase transition a local minimum must develop as the coupling is increased. This becomes deeper as the coupling is increased further until it is below the minimum at zero and so becomes the global minimum when the coupling crosses the critical value. A continuous phase transition appears when the global minimum ‘rolls’ out of the origin. This implies that the minimum at the origin must flatten out and become unstable. To investigate first order phase transitions one must be able to analyze the effective potential at arbitrarily large values of the order parameter. Unfortunately the effective potential is not reliable at large values of the order parameter because it is not possible to make a reliable truncation of the series of diagrams in the effective potential for large order parameters. For a continuous phase transition, on the other hand, the order parameter would be arbitrarily small close enough to the transition point. Hence, the order parameter can be used as an expansion parameter, allowing one to make reliable truncations of the series of diagrams. For this reason we restrict ourselves to the investigation of continuous phase transitions. In a stability analysis for continuous phase transitions the one–instanton amplitude is irrelevant. The reason is as follows. The effective potential consists of terms with positive powers of the order parameter. Near the origin the nondynamical kinetic part of the potential is second order in the order parameter. Hence, to destabilize the potential at the origin the dynamical terms must be of second order or smaller. Terms with higher powers of the order parameter will become insignificant compared to the kinetic term when the order parameter approaches zero. Looking at the one–instanton diagram, shown in Figure 1a, one sees that this term always comes with as many powers of the order parameter as it takes to close off all the zeromode lines. For $`N_f`$ flavors there would be $`N_f`$ factors of the order parameter. This means that for $`N_f>2`$ the one–instanton term cannot compete with the kinetic term near the origin. The leading contribution to the dynamic part of the potential that is second order in the order parameter is the instanton–anti-instanton term with two mass insertions shown in Figure 1b. Therefore only this term is considered in our calculation. It turns out that this term does not have an infrared problem in our stability analysis. The main result of our stability analysis is presented in terms of an expression for arbitrary $`N_f`$ and $`N_c`$ from which one can compute the critical couplings for chiral symmetry breaking via instantons. We determine the critical numbers of flavors for $`N_c=3,4,5`$ and $`6`$, assuming a two-loop beta function. Numerical values for critical couplings are provided in the fixed point regions of these values of $`N_c`$, up to the respective critical numbers of flavors. These results indicate that the critical couplings for the formation of 2-point functions via instantons are much smaller than those for gauge exchange dynamics when the number of flavors becomes large. The increasing strength of the instanton dynamics can be understood in terms of the increase in the number of fermion lines of the instanton vertices. The instanton effective action formalism is briefly reviewed in Section II. We distinguish between the dynamical and nondynamical terms of the effective action. Section III addresses the nondynamical terms and the dynamical term is considered in Section IV. We present and discussed the results of this calculation in Section V and conclude with a short summary in Section VI. ## II Instanton effective action formalism The 2-point effective action in the instanton formalism of Reference is given by $$\mathrm{\Gamma }[S]=\mathrm{Tr}\left\{\mathrm{ln}\left(S^1\right)\right\}+\mathrm{Tr}\left\{\left(S_m^1S^1\right)S\right\}+W_{2PI}[S]$$ (4) where the fermion modal propagator, $`S_m`$, consists of different propagators for the different types of modes associated with an instanton background configuration. We parameterize the full fermion propagator $`S`$ by $$S(p)=\frac{1}{ip/\mathrm{\Sigma }(p)}=\frac{ip/\mathrm{\Sigma }(p)}{p^2+\mathrm{\Sigma }^2(p)}$$ (5) where $`\mathrm{\Sigma }(p)`$ is the dynamical mass function. One can classify the terms of the effective action as either dynamical or nondynamical. Dynamical terms are those that vanish when the coupling is set to zero, while the nondynamical terms are those that remain. The first two one-loop kinetic terms in (4) are nondynamical terms. They consist of one-loop diagrams that are only composed of the full fermion propagator $`S`$ and the fermion modal propagator, $`S_m`$. The dynamical term $`W_{2PI}[S]`$ in (4) is the sum of all 2PI vacuum diagrams constructed with Feynman rules for: * the $`2N_f`$-point instanton vertices, $`𝒱_I`$ and $`𝒱_A`$, given below; * the full fermion propagator $`S`$, given in (5); and * an integral over the collective coordinates (sizes, positions and color orientations) of all instantons in the diagram. The dynamics in this analysis are provided by the instantons and are represented by the instanton vertices: $$𝒱_I=𝒜\underset{n}{\overset{N_f}{}}\left(\delta _0S_0^1\psi _n^0\right)\left(\overline{\psi }_n^0S_0^1\overline{\delta }_0\right),$$ (6) where $`S_0^1`$ is the inverse bare propagator and the functional derivatives $`\delta _0(=\delta /\delta \eta _0)`$ and $`\overline{\delta }_0(=\delta /\delta \overline{\eta }_0)`$ operate on the zeromode parts of the fermion sources, $`\eta _0`$ and $`\overline{\eta }_0`$. The expression for anti-instantons, $`𝒱_A`$, differs from the expression in (6) only in that the zeromode functions, $`\psi ^0`$ and $`\overline{\psi }^0`$, have the opposite helicities. The coefficient of the instanton vertex is the integrand of the ’t Hooft instanton amplitude : $$𝒜=\kappa \gamma (\alpha )\frac{1}{\rho ^5}(\mu \rho )^{b_0+N_f}.$$ (7) Here $`\rho `$ is a scale parameter for the size of the instanton and $`b_0=\frac{1}{3}(11N_c2N_f)`$. The dependence on the gauge coupling in (7) is contained in $$\gamma \left(\alpha (\mu )\right)=\left(\frac{2\pi }{\alpha (\mu )}\right)^{2N_c}\mathrm{exp}\left(\frac{2\pi }{\alpha (\mu )}\right).$$ (8) The factor $`(\mu \rho )^{b_0}`$ appearing in (7) can be incorporated into the exponent part of $`\gamma (\alpha )`$. This then leads to one-loop running as a function of $`\rho `$ for the coupling in the exponent in $`\gamma (\alpha )`$. The instanton expression in (7) can be computed to higher orders which would cause the coupling in the monomial part of $`\gamma (\alpha )`$ to run as a function of $`\rho `$ as well, but always at one order less than the coupling in the exponent. To all orders one would expect that both couplings run according to the exact beta function. Hence, assuming that the two-loop beta function is an accurate enough representation of the exact beta function, one can allow both couplings in $`\gamma (\alpha )`$ to run according to the two-loop beta function . This is what we shall assume so that we can make the replacement $`\gamma (\alpha (\mu ))(\mu \rho )^{b_0}\gamma (\alpha (\rho ))`$. The shape of $`\gamma (\alpha )`$ as a function of $`\alpha `$ is shown in Figure 2. The instanton amplitude peaks at a value of $`\alpha =\pi /N_c`$. The prefactor in (7) is $$\kappa =\frac{2\mathrm{exp}(5/6)\mathrm{exp}(BN_fCN_c])}{\pi ^2(N_c1)!(N_c2)!}$$ (9) where we follow and use the one-loop values: $`B=0.2917`$, $`C=1.5114`$ in MS-bar . Note that unlike the case in Reference the integration over the color orientations has not yet been done in (6). To alleviate the combinatorical calculations, this step is postponed until after the propagators are connected and unnecessary gauge transformations are cancelled. However, $`\kappa `$ in (9) already contains the volume of the factor group, $`SU(N_c)/G`$, where $`G`$ is the subgroup that leaves the instantons invariant. All that remains of the integration over color orientations is a color averaging integral. The expression of the fermion zeromode function in singular gauge that appears in (6) is $$\psi ^0(x)=\frac{\sqrt{2}\rho x_\mu \gamma _\mu U^\alpha }{\pi \sqrt{\mu }\sqrt{x^2}(\rho ^2+x^2)^{3/2}},$$ (10) where $`\mu `$ is the renormalization scale<sup>*</sup><sup>*</sup>*The $`\mu `$ is included here to get a mass dimension of $`\frac{3}{2}`$ for the fermion zeromodes. and $`x^2`$ denotes an Euclidean dot-product: $`x_\mu x_\mu `$. The Euclidean Dirac matrices, $`\gamma _\mu `$, are here defined so that $`\{\gamma _\mu ,\gamma _\nu \}=2\delta _{\mu \nu }\mathrm{𝟏}`$, $`\mathrm{tr}\{\gamma _\mu \gamma _\nu \}=4\delta _{\mu \nu }`$ and $`\gamma _\mu ^{}=\gamma _\mu `$. The $`U^\alpha `$ denotes a Dirac spinor with isospin index $`\alpha `$. (The isospin index is implicit in $`\psi ^0(x)`$.) This spinor is normalized such that $`_\alpha \overline{U}_\alpha U^\alpha =1`$ and it obeys the identity $`_\alpha U^\alpha \overline{U}_\alpha =\frac{1}{4}(1+\gamma _5)`$. We shall also later need the Fourier-transform of the zeromode function $$\mathrm{\Psi }^0(p)=\psi ^0(x)\mathrm{exp}(ixp)d^4x=i\frac{2\pi \sqrt{2}\rho p_\mu \gamma _\mu U^\alpha }{\sqrt{\mu }p^2}f\left(\frac{\rho \sqrt{p^2}}{2}\right),$$ (11) where the function, $$f(x)=2x\left[I_0(x)K_1(x)I_1(x)K_0(x)\right]2I_1(x)K_1(x),$$ (12) is defined in terms of modified Bessel functions . ## III The nondynamical terms The trace over the inverse modal propagator in (4) contains an integral over all the collective coordinates involved in the diagram. The general form of the result of this integration is the inverse bare propagator, $`S_0^1=ip/`$, multiplied by a momentum-dependent scalar form factor, which we shall assume to be equal to 1. In terms of (5) the nondynamical terms in (4), which we denote by $`\mathrm{\Gamma }_{kin}`$, becomes $$\mathrm{\Gamma }_{kin}=\left(\frac{4\mathrm{\Sigma }^2(p)}{p^2+\mathrm{\Sigma }^2(p)}2\mathrm{ln}\left[p^2+\mathrm{\Sigma }^2(p)\right]\right)\frac{d^4p}{(2\pi )^4}.$$ (13) It is customary to introduce an infrared cutoff at the momentum scale where $`p_0=\mathrm{\Sigma }(p_0)`$. Close to the phase transition point $`\mathrm{\Sigma }(p)`$ is very small so that $`p_00`$. An expansion of $`\mathrm{\Gamma }_{kin}`$ in term of $`\mathrm{\Sigma }(p)/p`$ contains only even powers. We drop the terms that are independent of $`\mathrm{\Sigma }(p)`$ because they do not affect the location of the minimum and we drop the terms that are higher than second order in $`\mathrm{\Sigma }(p)`$ because they are suppressed. The result is $$\mathrm{\Gamma }_{kin}^{(2)}=\frac{N_cN_f}{4\pi ^2}_0^{\mathrm{}}\mathrm{\Sigma }(p)^2p𝑑p.$$ (14) Now we make a redefinition of the integration variables and do a Fourier transformation. For this purpose we define $$p=\mu \mathrm{exp}(t)\mathrm{and}\mathrm{\Sigma }(p)=\mu \mathrm{exp}(t)_{\mathrm{}}^{\mathrm{}}\sigma (\omega )\mathrm{exp}(i\omega t)𝑑\omega $$ (15) where $`\mu `$ is the renormalization scale. Now we have $`\mathrm{\Gamma }_{kin}^{(2)}`$ $`=`$ $`{\displaystyle \frac{N_cN_f\mu ^4}{4\pi ^2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega _1)\sigma (\omega _2){\displaystyle _{\mathrm{}}^{\mathrm{}}}\mathrm{exp}(i\omega _1t)\mathrm{exp}(i\omega _2t)𝑑t𝑑\omega _1𝑑\omega _2`$ (16) $`=`$ $`{\displaystyle \frac{N_cN_f\mu ^4}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega )\sigma (\omega )𝑑\omega .`$ (17) ## IV The dynamical term The instanton–anti-instanton term that is quadratic in the dynamical mass looks as follows: each mass insertion closes off two lines of an instanton vertex. The remaining lines are all connected between the two vertices. (See Figure 1b.) The resulting diagram is expressed as $$W_{2PI}^{(2)}=\kappa ^2\gamma (\alpha (\rho _1))\gamma (\alpha (\rho _2))(\mu ^2\rho _1\rho _2)^{N_f}D(\rho _1)D(\rho _2)G(\rho _1,\rho _2)\frac{1}{\rho _1^5}\frac{1}{\rho _2^5}𝑑\rho _1𝑑\rho _2$$ (18) where $`\rho _1`$ and $`\rho _2`$ denote the sizes of the two instantons; the combinatorial factor $``$ simply counts the number of terms associated with this diagram; $`D(\rho _1)`$ and $`D(\rho _2)`$ are the parts of the expression due to the two mass loops and $`G(\rho _1,\rho _2)`$ is the part of the expression due to the remaining fermion loops. ### A The combinatorics and color averaging To count the number of terms we start by considering all the possible ways that the propagators can close off the lines of the two instanton vertices to give a diagram with two mass loops. Then we perform the color averaging and add up all the different terms. Recall that each line of an instanton vertex has a specific flavor associated with it. There is one incoming and one outgoing line for each flavor. Because the propagator is flavor diagonal it must connect lines of the same flavor. Hence, there are $`N_f`$ distinct ways to connect one of the two mass insertions, each leaving the other mass insertion and the remaining fermion lines with only one way to be connected. Next we perform the averaging over the color orientations of the two respective instantons. The color orientations are represented by gauge transformations that operate on the zeromodes: $$\psi ^0g\psi ^0\mathrm{and}\overline{\psi }^0\overline{\psi }^0g^1.$$ (19) In the mass loops they cancel because there we have a $`g`$ and a $`g^1`$ from the same instanton. The gauge transformations of the remaining lines can be combined into relative gauge transformations of one instanton with respect to the other: $`g_1^1g_2=g_r`$ or $`g_2^1g_1=g_r^1`$. Thus the two color averaging integrals become one trivial ($`=1`$) and one nontrivial averaging integral. The nontrivial integral contains $`(N_f1)`$ factors of $`g_r`$ and $`g_r^1`$. The result of the averaging is a series of tensor structures each multiplied by a coefficient that depends on $`N_f`$ and $`N_c`$. General expressions for these coefficients are provided in Reference . The tensor structures indicate how color indices of the same instanton are interconnected. This in turn determines how the fermion lines between the two instantons are closed off into fermion loops and thereby determines the form of the momentum integrals. The evaluation of all these momentum integrals is too challenging to attempt. However, we expect that the sizes of the momentum integrals do not differ by more that an $`O(1)`$ factor and that the variations would average out. Therefore we set them all equal and estimate their typical size in Section IV C. Each tensor structure consists of $`(N_f1)!`$ terms because of all the possible permutations of the $`(N_f1)`$ pairs of fermion lines between the two instantons. The different tensor structures denote different relative permutations of one line with respect to the other in the $`(N_f1)`$ pairs of fermion lines. All tensor structures that are associated with the same conjugacy class in the permutation group have the same coefficient. Because we assume that the momentum integrals are the same, we only have to add all the coefficients of all the tensor structures multiplied by the number of terms associated with each tensor structure. It can be shown that for permutations of $`(N_f1)`$ elements, $$\underset{classes}{}\alpha _iN_i=\frac{(N_c1)!}{(N_c+N_f2)!},$$ (20) where $`N_i`$ is the number of group elements in the $`i`$-th conjugacy class of the permutation group and $`\alpha _i`$ is the coefficient of the tensor structures in that class. To get the total count we have to multiply (20) by $`N_f`$ for the number of ways to connect the mass loops and by $`(N_f1)!`$ for the number of terms in each tensor structure. The resulting combinatorial factor is $$=\frac{N_f!(N_c1)!}{(N_c+N_f2)!}.$$ (21) ### B The mass insertions Using (5) and (11) we obtain the following expressions for the mass loops $`D(\rho )`$ $`=`$ $`{\displaystyle \overline{\mathrm{\Psi }}^0(p)S_0^1(p)S(p)S_0^1(p)\mathrm{\Psi }^0(p)\frac{d^4p}{(2\pi )^4}}`$ (22) $`=`$ $`{\displaystyle \frac{\rho ^2}{\mu }}{\displaystyle \frac{\mathrm{\Sigma }(p)}{p^2+\mathrm{\Sigma }^2(p)}f^2\left(\frac{\rho p}{2}\right)p^3𝑑p}`$ (23) $`=`$ $`{\displaystyle \frac{\rho ^2}{\mu }}{\displaystyle \mathrm{\Sigma }(p)f^2\left(\frac{\rho p}{2}\right)p𝑑p}+O\left(\mathrm{\Sigma }^3(p)/p^3\right)`$ (24) where we again made use of the approximations introduced in Section III. Next we again make a redefinition of the integration variables and do a Fourier transformation as in (15). In addition we define $$\frac{\rho \mu }{2}=\mathrm{exp}(\eta )\mathrm{and}f\left(\mathrm{exp}(a)\right)=f_0(a)$$ (25) As a result we have $`D(\rho )`$ $`=`$ $`4\mathrm{exp}(2\eta ){\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega )\mathrm{exp}(i\omega t)f_0^2(t+\eta )\mathrm{exp}(t)𝑑t𝑑\omega `$ (26) $`=`$ $`4\mathrm{exp}(\eta ){\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega )\mathrm{exp}(i\omega \eta )F(\omega )𝑑\omega `$ (27) where $$F(\omega )=_{\mathrm{}}^{\mathrm{}}\mathrm{exp}(i\omega a)f_0^2(a)\mathrm{exp}(a)𝑑a.$$ (28) ### C The remaining fermion loops The rest of the fermion lines between the two instanton vertices combine into $$G(\rho _1,\rho _2)=\underset{i}{\overset{2N_f2}{}}\left[\frac{d^4p_i}{(2\pi )^4}\overline{\mathrm{\Psi }}^0(p_i)S_0^1(p_i)S(p_i)S_0^1(p_i)\mathrm{\Psi }^0(p_i)\right](2\pi )^4\delta ^4\left(\underset{i}{}p_i\right).$$ (29) The fermion lines pair up into fermion loops due to the color contractions that appear after the averaging over color orientations. Considering one such loop, we find $`I(\rho _1,\rho _2,p,q)`$ $`=`$ $`\overline{\mathrm{\Psi }}_a^0(p)S_0^1(p)S(p)S_0^1(p)\mathrm{\Psi }_b^0(p)\overline{\mathrm{\Psi }}_b^0(q)S_0^1(q)S(q)S_0^1(q)\mathrm{\Psi }_a^0(q)`$ (30) $`=`$ $`{\displaystyle \frac{2(2\pi )^4}{\mu ^2}}\rho _1^2\rho _2^2{\displaystyle \frac{pq}{p^2q^2}}f\left({\displaystyle \frac{\rho _1p}{2}}\right)f\left({\displaystyle \frac{\rho _2p}{2}}\right)f\left({\displaystyle \frac{\rho _1q}{2}}\right)f\left({\displaystyle \frac{\rho _2q}{2}}\right)+O\left(\mathrm{\Sigma }_0^2/\mu ^2\right)`$ (31) where in the first line we explicitly show the color indices on the zeromodes. By defining $$J(k,\rho _1,\rho _2)=I(\rho _1,\rho _2,p,pk)\frac{d^4p}{(2\pi )^4},$$ (32) one can write (29) as $$G(\rho _1,\rho _2)=\underset{i}{\overset{N_f1}{}}\left[\frac{d^4k_i}{(2\pi )^4}J(k_i,\rho _1,\rho _2)\right](2\pi )^4\delta ^4\left(\underset{i}{}k_i\right).$$ (33) We shall not attempt to evaluate (33) exactly, but instead estimate its size in terms of the number of fermion loops. It turns out that the $`\rho `$-integrals are dominated in the region where $`\rho _1\rho _2`$. Therefore we set $`\rho _1=\rho _2=\rho `$ in the arguments of the $`f`$-functions. Then (32) becomes $$J(k,\rho _1,\rho _2)=\frac{8\pi \rho _1^2\rho _2^2}{\mu ^2}\frac{p^2pk}{p^2(pk)^2}f^2\left(\frac{\rho p}{2}\right)f^2\left(\frac{\rho \sqrt{(pk)^2}}{2}\right)p^3\mathrm{sin}^2(\theta )𝑑\theta 𝑑p.$$ (34) For the region where $`p>k`$ we set $`\sqrt{(pk)^2}=p`$ inside the $`f`$-function and for $`p<k`$ we set $`\sqrt{(pk)^2}=k`$. Upon evaluating the angular integral we arrive at $$J(k,\rho _1,\rho _2)=\frac{(2\pi )^2\rho _1^2\rho _2^2}{\mu ^2}\left[_0^k\frac{p^2}{2k^2}f^2\left(\frac{\rho p}{2}\right)f^2\left(\frac{\rho k}{2}\right)p𝑑p+_k^{\mathrm{}}\left(1\frac{k^2}{2p^2}\right)f^4\left(\frac{\rho p}{2}\right)p𝑑p\right].$$ (35) If we assume that $`\rho _1>\rho _2`$ then $`\rho =\rho _1`$, because the $`f`$-function with the largest $`\rho `$ dominates. Defining $$x=\frac{p\rho _1}{2}\mathrm{and}y=\frac{k\rho _1}{2}$$ (36) one can write (35) as $$J(k,\rho _1,\rho _2)=\frac{\pi ^2\rho _2^2}{\mu ^2}\left[\frac{8f^2(y)}{y^2}_0^yf^2(x)x^3𝑑x+16_y^{\mathrm{}}f^4(x)x𝑑x8y^2_y^{\mathrm{}}\frac{1}{x}f^4(x)𝑑x\right]\frac{\pi ^2\rho _2^2}{\mu ^2}R(y).$$ (37) The size of the area under $`R(y)`$ is $`\frac{1}{2}`$. This function determines the momentum dependence under the remaining momentum integrals in (33). To assess the size of the remaining integrals we note that (33) has $`(N_f2)`$ integrals over $`k_i`$. This will give $`(N_f3)`$ angular integrals. (One can arrange it that each $`k_i`$ at most contracts with its two neighbors). Together with the definitions in (36) we estimate the size of the remaining momentum integrals to be $$\underset{i}{\overset{N_f1}{}}\left[\frac{d^4k_i}{(2\pi )^4}R\left(\frac{k_i\rho _1}{2}\right)\right](2\pi )^4\delta ^4\left(\underset{i}{}k_i\right)\left(\frac{1}{2}\right)^{N_f1}\left[\frac{16}{\rho _1^4}(2\pi )^2\right]^{N_f2}.$$ (38) Our estimate for the sizes of $`G(\rho _1,\rho _2)`$ then gives $$G(\rho _1,\rho _2)\frac{\rho _1^4\pi ^2}{8}\left[\frac{2\rho _2^2}{\mu ^2\rho _1^4}\right]^{N_f1}.$$ (39) ## V Results and discussion The integrand of (18) is invariant under an interchange of $`\rho _1`$ and $`\rho _2`$. Therefore, one can split the $`\rho _2`$-integration into two regions, divided by the value of $`\rho _1`$. First, we consider only the part where $`\rho _2<\rho _1`$. When we substitute (39) into this part we find $$W_{\rho _2<\rho _1}^{(2)}=\frac{2^{N_f}\pi ^2}{16}\kappa ^2\mu ^2\gamma (\alpha (\rho _1))\gamma (\alpha (\rho _2))D(\rho _1)D(\rho _2)\rho _1^{33N_f}\rho _2^{3N_f7}𝑑\rho _1𝑑\rho _2.$$ (40) One can see that the $`\rho _2`$-integral is dominated by the region near the $`\rho _1`$-boundary. For this reason one can set $`\gamma (\alpha (\rho _2))=\gamma (\alpha (\rho _1))`$. The integrand of the $`\rho _1`$-integral has its dominant region around some value $`\rho _m`$, close to $`\mu `$. Since we consider cases with large numbers of flavors, the coupling runs slowly. Therefore one can set $`\gamma (\alpha (\rho _1))=\gamma (\alpha (\rho _m))=\gamma _0`$. Next we use (25) to redefine the variables in (40) and we substitute in (27). The result is $`W_{\rho _2<\rho _1}^{(2)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mu ^4\kappa ^2\pi ^22^{N_f}\gamma _0^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega _1)\sigma (\omega _2)F(\omega _1)F(\omega _2)`$ (42) $`\times {\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\eta _1}}\mathrm{exp}([53N_f+i\omega _1]\eta _1)\mathrm{exp}([3N_f5+i\omega _2]\eta _2)d\eta _2d\eta _1d\omega _1d\omega _2`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mu ^4\kappa ^2\pi ^32^{N_f}\gamma _0^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega _1)\sigma (\omega _2)F(\omega _1)F(\omega _2){\displaystyle \frac{\delta \left(\omega _1+\omega _2\right)}{3N_f5+i\omega _2}}𝑑\omega _1𝑑\omega _2`$ (43) Now we add the part where $`\rho _2>\rho _1`$. This is the same as (43) but with $`\omega _1`$ and $`\omega _2`$ interchanged. Then we evaluate one of the $`\omega `$-integrals. The result is $`W_{2PI}^{(2)}`$ $`=`$ $`W_{\rho _2<\rho _1}^{(2)}+W_{\rho _2>\rho _1}^{(2)}`$ (44) $`=`$ $`\mu ^4\kappa ^2\pi ^32^{N_f}\gamma _0^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}\sigma (\omega )\sigma (\omega )\left[{\displaystyle \frac{(3N_f5)F(\omega )F(\omega )}{(3N_f5)^2+\omega ^2}}\right]𝑑\omega .`$ (45) The maximum value of the part of the integrand in the brackets is at $`\omega =0`$ with $`F(0)^20.3`$. We notice that this term does not have an infrared divergence. Comparing the result in (45) with the nondynamical term in (17) one can see that the potential becomes unstable at the origin when $$\kappa ^2\pi ^32^{N_f}\gamma _0^2\left[\frac{F(0)^2}{3N_f5}\right]>\frac{N_cN_f}{2\pi }.$$ (46) This can be expressed as $$\gamma _0>1,$$ (47) where $$(N_c,N_f)=\frac{2F(0)\mathrm{exp}(5/6)\mathrm{exp}(BN_fCN_c)}{(N_c2)!}\sqrt{\frac{2^{N_f+1}(N_f1)!}{(3N_f5)(N_f+N_c2)!N_c!}}$$ (48) only depends on $`N_c`$ and $`N_f`$. The critical coupling for instanton dynamics $`\alpha _c^I`$ is the one for which $$\gamma \left(\alpha (\rho _m)=\alpha _c^I\right)=\frac{1}{(N_c,N_f)}.$$ (49) It necessarily has to be smaller than $`\pi /N_c`$. The requirement in (49), together with (48), is the main result of our analysis. From it one can compute $`\alpha _c^I`$ for any value of $`N_c`$ and $`N_f>4/3`$. Numerical estimates of $`\alpha _c^I`$ for various values of $`N_c`$ and $`N_f`$ are provided in Figure 3. Before we discuss these critical couplings we need to clarify the ranges of numbers of flavors for which these critical couplings are computed. These ranges are defined under the assumption that the running of the coupling is governed by the two-loop beta function – an assumption which becomes more accurate as the coupling becomes smaller. We only compute the critical couplings in the region of $`N_f`$ where the two-loop beta function has an infrared fixed point. The left-hand end of each curve in Figure 3 is therefore the point where a fixed point value appears, which moves down from infinity as a function of increasing $`N_f`$. The right-hand ends of the curves in Figure 3 indicate the critical numbers of flavors $`N_f^c`$ above which chiral symmetry breaking does not occur. When there is a nontrivial infrared fixed point in the beta function the maximum coupling that can be reached is the fixed point coupling. Assuming a two-loop beta function the fixed point coupling is given by $$\alpha _{}=\frac{b}{c}=\frac{4\pi (11N_c2N_f)N_c}{34N_c^313N_c^2N_f+3N_f}$$ (50) where $`b`$ and $`c`$ are respectively the first and second coefficients in the beta function. The critical numbers of flavors $`N_f^c`$ is reached when the fixed point coupling given by (50) falls below the critical coupling, obtained from solving (49). Therefore for values of $`N_f`$ above $`N_f^c`$ the coupling is unable to reach the critical value so that chiral symmetry breaking does not occur. The critical number of flavors, $`N_f^c`$ are $`13.48,17.95,22.41`$ and $`26.87`$ for $`N_c=3,4,5,`$ and $`6`$, respectively. This gives $`N_f^c4.5N_c`$, compared to $`4.77N_c`$, reported in Reference . We computed critical couplings for $`N_c=3,4,5`$ and $`6`$ and for $`N_f`$ in the fixed point region up to $`N_f^c`$. The results are shown in Figure 3. From these results we make the following observations: * The critical coupling decreases as either $`N_c`$ or $`N_f`$ increases. * The values of the critical couplings near $`N_f^c`$ are small compared to the equivalent critical coupling for gauge exchanges given in (2). Compare, for example, the $`N_c=3`$ values. We obtain $`\alpha _c^I=0.37`$, while the equivalent gauge exchange value is $`\alpha _c^{ge}=\pi /4=0.785`$. * Toward smaller values of $`N_f`$ the instanton critical couplings become more comparable to their gauge exchange counterparts. On this basis we conclude that instanton dynamics are much stronger than gauge dynamics for large numbers of flavors. ## VI summary We have used the instanton effective action formalism of Reference to compute the critical couplings for the formation of 2-point functions. For this analysis we restricted ourselves to continuous phase transitions for chiral symmetry breaking, which implies that we only consider the instanton–anti-instanton amplitude with two mass insertions. We found that this amplitude does not suffer from an infrared divergence. Couplings are allowed to run according to the two-loop beta functions. The resulting critical couplings indicate that instanton dynamics are much stronger than gauge dynamics for large numbers of flavors. ## Acknowledgements The author wants to express his appreciation for the valuable comments of Bob Holdom. He also wants to thank Craig Burrell for his help in the preparation of this manuscript.
warning/0003/cond-mat0003081.html
ar5iv
text
# Bose condensates at high angular momenta ## Abstract We exploit the analogy with the Quantum Hall (QH) system to study weakly interacting bosons in a harmonic trap. For a $`\delta `$-function interaction potential the “yrast” states with $`LN(N1)`$ are degenerate, and we show how this can be understood in terms of Haldane exclusion statistics. We present spectra for 4 and 8 particles obtained by numerical and algebraic methods, and demonstrate how a more general hard-core potential lifts the degeneracies on the yrast line. The exact wavefunctions for $`N=4`$ are compared with trial states constructed from composite fermions (CF), and the possibility of using CF-states to study the low $`L`$ region at high $`N`$ is discussed. PACS numbers: 03.75.Fi, 05.30.Jp, 73.40.Hm The close relation between high angular momentum states of a condensate of weakly interacting hard core bosons and the Quantum Hall (QH) effect was recently pointed out . The essential observation is that the weak interaction limit allows for a two-dimensional description of the boson system in terms of lowest Landau level (LLL) wave functions – just as for a QH system, and they are both described by wave functions containing powers of the Laughlin-Jastrow factor $`_{i<j}(z_iz_j)`$, where $`z_i`$ is the complex coordinate of the $`i^{th}`$ particle. As pointed out by Cooper and Wilkin , the two systems can in fact be mapped onto each other by a standard Leinaas-Myrheim transformation , attaching an odd number of units of statistical flux to each particle. This enables us to use both intuition and techniques from the QH system to study rotating Bose condensates. In this paper we shall mainly discuss the angular momentum region $`N(N1)L2N(N1)`$, where the ground state corresponding to a $`\delta `$-function two-body interaction is degenerate. We explicitly show how these degeneracies can be understood via a mapping to a system of free anyons in the LLL, and then show how the degeneracy is broken by a more general short range potential containing derivatives of delta functions. We also make a detailed comparison between algebraically calculated exact wave functions and trial wave functions formed from so called compact states of composite fermions, a construction orginally due to Jain and Kawamura . In particular, we will emphasize the importance of a certain class of wavefunctions where the polynomial part is translationally invariant. Although for computational reasons we have results only for few particles, $`N=4,6`$ and 8, it is clear that some of our results, like the degeneracy structure of the yrast line, hold for any $`N`$. We also believe that many features of the results for the CF wave functions will generalize to higher $`N`$ . Since the flux attachment changes the angular momentum by $`mN(N1)/2`$, where $`m`$ is the number of fluxes attached, the boson - fermion mapping would apparently only be useful for studying angular momenta that are out of reach of present experiments (which are limited to the strong interaction regime and $`LN`$). However, there are some indications that fermionic techniqes could be useful for $`L`$ as low as $`N`$, i.e. for the so called single vortex state. Although we will return to this point at the end of the paper, we shall for now, without any further apologies, consider the theoretical problem of understanding the region $`N(N1)L2N(N1)`$ of the yrast line. The simplest model for a hard-core interaction is a delta function potential. We thus consider a model of $`N`$ interacting spinless bosons in a harmonic trap of strength $`\omega `$. In the limit of weak interaction, this may be rewritten as a two-dimensional lowest Landau level (LLL) problem in the effective “magnetic” field $`B_{eff}=2m\omega `$ with the Hamiltonian taking the form $`H=\omega L+g{\displaystyle \underset{i<j}{}}\delta ^2(𝐫_i𝐫_j)`$ (1) ($`\mathrm{}=1`$), where $`L_il_i=L_z`$ is the total angular momentum. The single-particle states spanning our Hilbert space are $`\eta _{0,l}=(2^{l+1}\pi l!)^{1/2}z^l\mathrm{exp}(\overline{z}z/4)`$ with $`z=\sqrt{2m\omega }(x+iy)`$. In Figs. 1 and 2 we show the interaction energy, in units of $`g/4\pi `$ as a function of the total angular momentum $`L`$ for $`N=4`$, $`L20`$ and $`N=8`$, $`30L56`$, respectively. The many-body states are obtained from Lanczos diagonalization suitable for large and relatively sparse matrices . The Fock space is spanned by single-particle states that are characterized only by the positive quantum numbers $`l`$, where $`0lL`$. (Similar exact diagonalization studies have recently been reported in and .) We note the following properties in the many-body spectra: Since increasing the angular momentum spreads out the particles in space, the yrast energy, i.e. the lowest possible interaction energy for given $`L`$, decreases with increasing $`L`$. For each state in the spectrum, there exists a set of “daughter states” with higher values of $`L`$, having exactly the same (interaction) energy as the original state. These daughters are simply center-of-mass excitations of the original states, thus having the same many-body correlations. For $`LN(N1)`$ there are zero energy states, which can be understood by noting that any wavefunction of the form $`\psi (z_1,z_2,\mathrm{},z_N)={\displaystyle \underset{i<j}{}}(z_iz_j)^2S(z_1,z_2,\mathrm{},z_N),`$ (2) where $`S(z_1,z_2,\mathrm{},z_N)`$ is a symmetric polynomial in the $`z_i`$:s, has zero interaction energy. Since the factor $`_{i<j}(z_iz_j)^2`$ contributes an angular momentum $`L_0=N(N1)`$, states of the type (2) exist for $`LL_0`$. At $`L=N(N1)`$ there is a unique state with zero interaction energy corresponding to $`S(z_1,z_2,\mathrm{},z_N)=1`$, while the states at higher $`L`$ typically are degenerate. The systematics of these degeneracies can be understood by a mapping to anyons in the lowest Landau level. The essential observation is that the wave functions (2) describe anyons in the LLL with statistics parameter $`\alpha =2`$ (in general, the statistics parameter is given by the exponent of the Jastrow factor). It is known that anyons in the LLL obey Haldane’s fractional exclusion statistics (FES), and following Ref., one can use this knowledge to construct the allowed many-body states for given $`N`$ and $`L`$ as angular momentum excitations of the Laughlin-like state at $`L=N(N1)`$. According to the definition of FES, each particle in the system blocks $`\alpha =2`$ single-particle states, and many-particle states with total angular momentum $`L`$ are constructed by occupying single-particle states, with a minimum distance of $`\alpha =2`$ between each pair of occupied levels (for an example, see Fig. 3). The number of allowed configurations then gives the degeneracy of the state for a given $`L`$. Table 1 shows the degeneracies of some of the states on the $`N=4`$ yrast line, as obtained from the anyon mapping, and they are in exact agreement with our numerical results. This construction implicitly uses that all the eigenstates on the yrast line for $`L>N(N1)`$ contain the Jastrow factor $`_{i<j}(z_iz_j)^2`$, so the degeneracies can also be found as the number of ways one can distribute $`M=LL_0`$ units of angular momentum among $`N`$ particles. For a more general hard core potential, the $`E=0`$ states on the yrast line above $`L=N(N1)`$ are no longer degenerate. To demonstrate this point and study how the degeneracy is broken, we add a potential of the form $`V=c_1^2\delta ^2(zz^{})+c_2^4\delta ^2(zz^{}),`$ (3) that was originally used by Trugman and Kivelson in the context of the fractional QH effect. The term $`^2\delta ^2(zz^{})`$ does not contribute to the interaction energy for fully symmetric states, whereas the $`^4\delta ^2(zz^{})`$ term gives small corrections to the spectra in Figs. 1 and 2 (at the percent level for the parameters used in the inset of Fig. 4, where we show regularized forms of the potentials (1) and (3)). We have examined how the potential (3) splits up the degeneracies of the zero interaction energy yrast states, by exact algebraic diagonalization, using computer algebra. Here we have directly used the form (2), and systematically exploited that for a given $`L`$, all states corresponding to center-of-mass excitations of lower $`L`$-eigenstates are orthogonal to the subspace of “new” states. The latter subspace consists of translation invariant (TI) polynomials, i.e. functions invariant under a simultaneous, constant shift $`z_iz_i+a`$ of all the coordinates. Following Trugman and Kivelson, we have used a basis constructed from elementary symmetric functions $`s_n`$. For given $`N`$ and $`L`$, the basis consists of all possible combinations $`|k_1k_2\mathrm{}k_ns_1^{k_1}(z_i)s_2^{k_2}(\stackrel{~}{z}_i)\mathrm{}s_N^{k_N}(\stackrel{~}{z}_N){\displaystyle \underset{i<j}{}}(z_iz_j)^2,`$ (4) such that $`_{n=1}^Nnk_n=LL_0`$. Note that, for $`n2`$, we have introduced the new variables $`\stackrel{~}{z}_iz_iz_c`$, with $`z_c`$ the center-of-mass coordinate $`z_c=(z_i)/N`$. The basis states spanning the TI subspace are identified as those with $`k_1=0`$. The diagonalization is thus performed within this subspace only, which reduces the matrix dimension substantially. The resulting spectrum for $`N=4`$, with the coefficient $`c_2`$ in (3) set equal to 1, is shown in Fig. 4. We notice the close similarity between Fig. 1 and 4. In both cases, the yrast line passes through the same number of steps, with the same step lengths, as $`L`$ is increased by $`N(N1)`$, to the point where the yrast energy becomes zero. At the point $`L=2N(N1)`$, the zero-energy yrast state is again non-degenerate and of the Laughlin type, i.e. $`_{i<j}(z_iz_j)^4`$, whereas the subsequent yrast states have degeneracies corresponding to FES with statistics parameter $`\alpha =4`$. Including even higher derivative terms in the repulsive potential would subsequently split up the higher regions of the yrast line. Finally, note that the yrast states corresponding to cusps, i.e. states that are followed by a “plateau” in the yrast line, are always in the TI subspace. The algebraic diagonalization used here is limited in practice to smaller particle numbers and angular momenta than the numerical scheme used in Figs. 1 and 2. However, the present approach has the advantage that it provides explicit, analytic expressions for the eigenfunctions, in terms of symmetric polynomials. This gives some additional insight into the structure of the yrast states, and in particular the region below the single vortex in the case of a pure delta function interaction. Bertsch and Papenbrock recently proposed and numerically tested the following form for the yrast states at $`2LN`$, $`\psi _L(z_i)`$ $`=`$ $`{\displaystyle \underset{p_1<p_2<\mathrm{}<p_L}{}}(z_{p_1}z_c)(z_{p_2}z_c)\mathrm{}(z_{p_L}z_c).`$ (5) This is just the symmetric polynomial $`s_L(\stackrel{~}{z}_i)`$, i.e. the state $`|0\mathrm{}\mathrm{1..0}`$ (with the 1 in the $`L`$th place), in the notation of (4) (without the Jastrow factor in the present case of a pure delta function interaction). This state is a basis state in the TI subspace for all $`2LN`$. In the cases where this is the only basis state ($`L=2,3`$ for $`N3`$), it is thus obvious that (5) is exact. Furthermore, performing the algebraic diagonalization up to $`L=N`$ for $`N=4`$ and $`N=6`$, we have confirmed that even when the translation invariant subspace is spanned by more than one basis vector, the basis state (5) is always an exact eigenstate. We now turn to a study of a class of wave functions that can be constructed in analogy with the so-called Jain states for the fractional QHE. The main idea is to map the strongly interacting LLL bosons to weakly interacting composite fermions by attaching an odd number $`m`$ of flux quanta to each particle. Trial wavefunctions with angular momentum $`L`$ are thus constructed as non-interacting fermionic wavefunctions with angular momentum $`Lm/2N(N1)`$, multiplied by $`m`$ Jastrow factors and projected onto the LLL, $`\psi _L=𝒫\left(f_F(z_i,\overline{z}_i){\displaystyle \underset{i<j}{}}(z_iz_j)^m\right).`$ (6) Here, $`f_F`$ is a Slater determinant consisting of single-particle wave functions $`\eta _{n,l}(z,\overline{z})z^lL_n^l(\overline{z}z/2)`$ where $`n`$ is the Landau level index ( $`ln`$), and $`L_n^l`$ a generalized Laguerre polynomial. Originally used for the homogeneous states relevant for the fractional QHE, wavefunctions of the type (6) were later employed to describe inhomogeneous systems such as quantum dots and recently by Cooper and Wilkin to study the bosonic yrast lines for up to 10 particles in the case of a pure delta function interaction. In short, the LLL projection in (6) amounts to the replacement $`\overline{z}_i2_i`$ in the polynomial part of the wave function. However, there are several ways of doing this in practice, and we shall compare the different projection methods when constructing trial wave functions for the yrast states in Figs. 1 and 4. The most straightforward way is to replace the $`\overline{z}`$:s with derivatives in the final polynomial, obtained after multiplying out the Slater determinant and the Jastrow factors and moving all $`\overline{z}`$:s to the left. In practice, this method is applicable only for small particle numbers, when the number of derivatives involved is not too large. We shall refer to this method as “Method I” and apply it in a few special cases for comparison. Alternatively, noting that $`\left|\begin{array}{ccc}\eta _{11}& \eta _{12}& ..\\ \eta _{21}& \eta _{22}& ..\\ \mathrm{}& \mathrm{}& ..\\ \eta _{N1}& \eta _{N2}& ..\end{array}\right|{\displaystyle \underset{i<j}{}}(z_iz_j)^{2p}=\left|\begin{array}{ccc}\stackrel{~}{\eta }_{11}& \stackrel{~}{\eta }_{12}& ..\\ \stackrel{~}{\eta }_{21}& \stackrel{~}{\eta }_{22}& ..\\ \mathrm{}& \mathrm{}& ..\\ \stackrel{~}{\eta }_{N1}& \stackrel{~}{\eta }_{N2}& ..\end{array}\right|,`$ (15) where $`\eta _{ij}\eta _i(z_j,\overline{z}_j)`$ and $`\stackrel{~}{\eta }_{ij}\eta _i(z_j,\overline{z}_j)_{kj}(z_jz_k)^p`$, one can first absorb $`2p`$ Jastrow factors in the Slater determinant, and then project entry by entry. Since the wave function (6) contains an odd number $`m`$ of Jastrow factors, one finally has to compensate by multiplying the resulting polynomial by $`_{i<j}(z_iz_j)^{m2p}`$. We shall refer to this as Method II and use it as follows; for a delta function interaction, where the wave function contains $`m=1`$ Jastrow factor, it is appropriate to use $`p=1`$ (Method IIa). In the case of the $`^4\delta `$-potential, the Jain construction will involve absorbing $`m=3`$ flux quanta in the wave function, and we shall compare projection with $`p=1`$ (Method IIa) and $`p=2`$ (Method IIb). (Note that Method IIa is in a sense trivial, since it relates the wavefunctions at $`L`$ and $`L+N(N1)`$ by multiplication with two Jastrow factors.) We have already stressed the significance of the TI subspace - these are the states that determine the shape of the yrast line. It is very appealing that there is a special set of the states (6) that are in this subspace, namely the so called compact states which are characterized by having the $`n`$:th Landau level occupied from $`l_n=n`$ to $`l_n=l_n^{max}`$ without any “holes”. In the context of the QH effect, the important property of the compact states is that they are homogeneous. When describing quantum dots using the non-interacting composite fermion model (NICFM), one can show that CF candidates for the cusp states must be compact, and the same line of arguments was later used for bosons . From the point of view of the wavefunctions, the importance of the compact states is that they are in the TI subspace, and it is rather remarkable that for all L where a compact state can be constructed, the one with the lowest CF energy has a large overlap with the lowest exact state in the TI subspace. This is true, independent of whether or not the state is a ground state, i.e. corresponds to a cusp in the yrast line. In the following discussion, we shall only be concerned with the wavefunctions, and will not discuss whether or not the NICFM can give a good description of the energy spectrum, a question that was already discussed in some detail . $`\delta `$-function potential: In this case we have calculated the Jain wavefunctions (6) corresponding to all compact states for $`N=4`$, $`0L12`$, taking $`m=1`$ and using projection method IIa ($`p=1`$). If there are two compact states with the same $`L`$, we use the one with the lowest CF effective energy. In table II we show the overlap with the exact algebraic wavefunctions. A “1” indicates that the wavefunctions are identical. For comparison we have also included the overlaps with the wavefunctions corresponding to projection method I, as given by Cooper and Wilkin. We note that all the compact states have very large overlap with the exact eigenstates. That the CF wavefunction is exact for $`L=2,3`$ is a simple consequence of the TI property and that there is only one state in the TI subspace at these $`L`$ values. That the $`L=7`$ state comes out exact is more surprising, and we have no good explanation for this. It would be interesting to pursue the exact diagonalization to higher $`N`$ in order to see if there are other non-trivial CF states that are exact. We also note that the direct projection (method I) does slightly better than method II. $`^4\delta `$-function potential: Here we calculated the Jain wavefunctions (6) corresponding to all, lowest CF-energy, compact states for $`N=4`$, $`12L24`$ taking $`m=3`$ and using projection methods IIa ($`p=1`$) and IIb $`(p=2`$). For comparison, we also used method I for three $`L`$ values. In table III we show the overlap with the exact algebraic wavefunctions. We again note that all the compact states (except $`L=17`$ with projection method I) have very large overlap with the exact eigenstates, and that for $`L=17,19`$, the CF state is not the ground state. Comparing the different projection methods, we see that in many, but not all cases they give identical results. In particular, one should note that method IIb does not reproduce the exact wavefunction for $`L=14,15`$ where the TI subspace is non-degenerate. This means that that projection gives a wave function that is not in the space of states given by (2). The same effect is even more striking for $`L=17`$, where method IIb gives the exact wave function, whereas direct projection (method I) gives a rather poor overlap, indicating that a large component is not in the subspace (2). Finally, we comment on the possibility to use CF wavefunctions in the experimentally more relevant case of low $`L`$, and in particular for the single vortex state at $`L=N`$. Rather surprisingly, the overlaps between the CF and the exact wavefunction (5) tend to get larger with increasing particle number, at least up to $`N=10`$. It thus seems worthwhile to use the CF approach to construct trial wavefunctions for the single vortex for general $`N`$. The single vortex CF state is in fact unique, if one demands that in addition to being compact, it should also have minimal CF cyclotron energy. The relevant Slater determinant is formed from the single particle states $`\eta _{n,n}`$ for $`n=N2,N1,\mathrm{}0`$ and $`\eta _{01}`$, and, using projection method I, the resulting wavefunction takes the following rather compact form, $`\psi _{L=N}={\displaystyle \underset{n=1}{\overset{N}{}}}(1)^nz_n{\displaystyle \underset{k<l;k,ln}{}}\left(_k_l\right){\displaystyle \underset{i<j}{}}(z_iz_j).`$ (16) Although the $`N^2`$ derivatives make it difficult to evaluate this function for large $`N`$, it can easily be handled in integrals of the form $`_id^2z_i\mathrm{exp}(1/2_i\overline{z}_iz_i)f(\overline{z}_i)\psi _{L=N}`$ by partial integrations. Alternatively one can use projection Method II, where the wavefunction becomes a determinant of linear combinations of elementary symmetric polynomials. In both cases, it should be possible to compare with the exact wavefunction (5) using Monte Carlo methods. ACKNOWLEDGEMENT: We are very grateful to B. Mottelson for introducing us to the physics of rotating Bose condensates, and for numerous fruitful discussions. We also wish to thank M. Manninen, G. Kavoulakis and R.K. Bhaduri for discussions and M. Koskinen for giving advice on the numerical work. This work was financially supported by the Academy of Finland, the Swedish Natural Science Research Council, the TMR programme of the European Community under contract ERBFMBICT972405, the “Bayerische Staatsministerium für Wissenschaft, Forschung und Kunst”, and the NORDITA Nordic project “Confined electronic systems”.
warning/0003/hep-th0003274.html
ar5iv
text
# Confined Maxwell Field and Temperature Inversion Symmetry ## 1 <br>Introduction. Temperature inversion symmetry occurs in the free energy associated with the Casimir effect at finite temperature and it is linked to the nature of the boundary conditions imposed on the quantum vacuum oscillations. As shown by Ravndal and Tollefsen , temperature inversion symmetry holds for massless bosonic fields and symmetric boundary conditions and also for massless fermionic fields and antisymmetric boundary conditions. Temperature inversion symmetry was initially studied by Brown and Maclay who wrote the scaled free energy associated with the Casimir effect at finite temperature as a sum of three contributions, namely: a zero temperature contribution, that is, the Casimir energy density at zero temperature, a Stefan-Boltzmann contribution proportional the fourth power of the scaled temperature $`\xi :=Td/\pi `$, and a non-trivial contribution. The three contributions to the scaled free energy can be combined into a single double sum: $$f(\xi )=\frac{1}{16\pi ^2}\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\underset{m=\mathrm{}}{\overset{\mathrm{}}{}}\frac{(2\pi \xi )^4}{[m^2+(2\pi n\xi )^2]^2},$$ (1) where the term corresponding to $`n=m=0`$ is excluded. Setting $`n=0`$ and summing over $`m`$ with $`m0`$, we obtain the Stefan-Boltzmann limit $`\pi ^6\xi ^4/45`$. On the other hand, setting $`m=0`$ and summing over $`n`$ with $`n0`$, we obtain the zero temperature Casimir term $`\pi ^2/720`$. This function has the following property: $$(2\pi \xi )^4f\left(1/4\pi \xi \right)=f\left(\xi \right),$$ (2) which is the mathematical statement of the temperature inversion symmetry. It was also shown by Gundersen and Ravndal that the scaled free energy associated with massless fermions fields at finite temperature submitted to MIT boundary conditions satisfy the relation given by Eq.(2) and therefore exhibts temperature inversion symmetry. Tadaki and Takagi have calculated Casimir free energies for a massless scalar field obeying Dirichlet or Neumann boundary conditions on both plates and found this symmetry. For the parallel plate geometry and mixed boundary conditions it is possible to circunvent the restrictions found by and discuss temperature inversion symmetry by reconizing that with respect to the evaluation of the free energy this arrangement is equivalent to the difference between two Dirichlet (or Neumann) plates . In the case of a massless scalar field at finite temperature and periodic boundary conditions, it is possible to show that the partition function, and consequently the free energy, can be written in a closed form such that the temperature inversion symmetry becomes explicit . Here we shall show that for the case of the Maxwell field confined by a perfectly conduting rectangular cavity this symmetry is also present, provided that we generalize the transformations that relate the high and the low temperature regimes. Throughout this letter we employ units such that Boltzmann constant $`k_B`$, the speed of light $`c`$ and $`\mathrm{}=h/2\pi `$ are set equal to the unity. ## 2 Evaluation of Helmholtz free energy Helmholtz free energy for the confined Maxwell field within a perfectly conducting rectangular box can be evaluated by means of the generalized zeta function technique , . First we introduce the global zeta function which is defined by $$\mu ^{2s}\zeta \left(s\right)=\mu ^{2s}\underset{p=\mathrm{}}{\overset{\mathrm{}}{}}\underset{\{n\}}{}\left[\left(\frac{2\pi p}{\beta }\right)^2+\omega _{\{n\}}^2\right]^s$$ (3) where $`\beta =1/T`$ , the reciprocal of the temperature $`T,`$ is the periodic length along the Euclidean time direction and a mass scale parameter $`\mu `$ was introduced in order to keep Eq.(3) dimensionless; $`\omega _{\{n\}}^2`$ are the eigenvalues associated with the Euclidean time-independent modal equation $$\mathrm{}_𝐱\phi _{\{n\}}(𝐱)=\omega _{\{n\}}^2\phi _{\{n\}}(𝐱),$$ (4) where $`𝐱=(x,y,z)`$ and $`\mathrm{}_𝐱`$ is the Laplacian operator. The Helmholtz free energy function $`F\left(\beta \right)`$ for bosonic fields can be obtained from the knowledge of the global zeta function through the relation $$F\left(\beta \right)=\frac{1}{2\beta }\frac{}{s}\left[\mu ^{2s}\zeta \left(s\right)\right]_{s=0}.$$ (5) The eigenfrequencies associated with the allowed electromagnetic normal modes within a perfectly conducting rectangular box of linear dimensions $`a,b,c`$ are given by $$\omega _{lmn}^2=\left(\frac{l\pi }{a}\right)^2+\left(\frac{m\pi }{b}\right)^2+\left(\frac{n\pi }{c}\right)^2,$$ (6) where $`l,m,n`$ $`\{1,2,\mathrm{}\}`$. For $`l,n,m0`$ there are two possible polarization states and for $`l=0`$ or $`m=0`$ or $`n=0`$ there is one polarization state only. Eigenfrequencies for which three or two of the integers $`l,m,n`$ are simultaneously zero are not allowed. It follows that the generalized zeta function for the case in question reads $`\mu ^{2s}\zeta \left(s\right)`$ $`=`$ $`2\mu ^{2s}{\displaystyle \underset{l,m,n=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{p=\mathrm{}}{\overset{\mathrm{}}{}}}\left[\left({\displaystyle \frac{l\pi }{a}}\right)^2+\left({\displaystyle \frac{m\pi }{b}}\right)^2+\left({\displaystyle \frac{n\pi }{c}}\right)^2+\left({\displaystyle \frac{2p\pi }{\beta }}\right)^2\right]^s`$ (7) $`+\mu ^{2s}{\displaystyle \underset{l,m=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{p=\mathrm{}}{\overset{\mathrm{}}{}}}\left[\left({\displaystyle \frac{l\pi }{a}}\right)^2+\left({\displaystyle \frac{m\pi }{b}}\right)^2+\left({\displaystyle \frac{2p\pi }{\beta }}\right)^2\right]^s`$ $`+\mu ^{2s}{\displaystyle \underset{l,n=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{p=\mathrm{}}{\overset{\mathrm{}}{}}}\left[\left({\displaystyle \frac{l\pi }{a}}\right)^2+\left({\displaystyle \frac{n\pi }{c}}\right)^2+\left({\displaystyle \frac{2p\pi }{\beta }}\right)^2\right]^s`$ $`+\mu ^{2s}{\displaystyle \underset{m,n=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{p=\mathrm{}}{\overset{\mathrm{}}{}}}\left[\left({\displaystyle \frac{m\pi }{b}}\right)^2+\left({\displaystyle \frac{n\pi }{c}}\right)^2+\left({\displaystyle \frac{2p\pi }{\beta }}\right)^2\right]^s,`$ If we now separate the terms corresponding to the zero temperature sector by setting $`p=0`$, then it will be easily seen that we can rewrite Eq.(7) formally as a sum of Epstein functions (see the Appendix) which reads $`2^2\left({\displaystyle \frac{\pi }{\mu }}\right)^{2s}\zeta (s)`$ $`=`$ $`2^4E_4(s;1/a^2,1/b^2,1/c^2,4/\beta ^2)+2^3E_3(s;1/a^2,1/b^2,1/c^2)`$ (8) $`+2^3E_3(s;1/a^2,1/b^2,4/\beta ^2)+2^2E_2(s;1/a^2,1/b^2)`$ $`+2^3E_3(s;1/a^2,1/c^2,4/\beta ^2)+2^2E_2(s;1/a^2,1/c^2)`$ $`+2^3E_3(s;1/b^2,1/c^2,4/\beta ^2)+2^2E_2(s;1/b^2,1/c^2)`$ In order to regularize Eq.(8) it is convenient to rewrite it as a sum of generalized zeta functions and this can be accomplished by making use of Eqs. ($`(A.4)`$) and (Appendix) in the appendix. Equation (8) then takes the form $`2^2\left({\displaystyle \frac{\pi }{\mu }}\right)^{2s}`$ $`\zeta `$ $`\left(s\right)=A_4(s;1/a^2,1/b^2,1/c^2,4/\beta ^2)A_2(s;1/a^2,4/\beta ^2)`$ (9) $`A_2(s;1/b^2,4/\beta ^2)A_2(s;1/c^2,4/\beta ^2)+4E_1(s;4/\beta ^2).`$ Finally, we can regularize Eq.(9) by applying the reflection formula, Eq.($`(A.3)`$) on all terms except the last one which is already regularized thus obtaining $`\mu ^{2s}\zeta \left(s\right)`$ $`=`$ $`{\displaystyle \frac{\mu ^{2s}abc\beta }{8\pi ^2}}{\displaystyle \frac{\mathrm{\Gamma }\left(2s\right)}{\mathrm{\Gamma }\left(s\right)}}A_4(2s;a^2,b^2,c^2,\beta ^2/4)`$ (10) $`{\displaystyle \frac{\mu ^{2s}a\beta }{8\pi ^2}}{\displaystyle \frac{\mathrm{\Gamma }\left(1s\right)}{\mathrm{\Gamma }\left(s\right)}}A_2(1s;a^2,\beta ^2/4)`$ $`{\displaystyle \frac{\mu ^{2s}b\beta }{8\pi ^2}}{\displaystyle \frac{\mathrm{\Gamma }\left(1s\right)}{\mathrm{\Gamma }\left(s\right)}}A_2(1s;b^2,\beta ^2/4)`$ $`{\displaystyle \frac{\mu ^{2s}c\beta }{8\pi ^2}}{\displaystyle \frac{\mathrm{\Gamma }\left(1s\right)}{\mathrm{\Gamma }\left(s\right)}}A_2(1s;c^2,\beta ^2/4)+\left({\displaystyle \frac{\beta }{2}}\right)^{2s}\zeta _R\left(2z\right),`$ where $`\zeta _R\left(2z\right)`$ is the Riemann zeta function and we have also made use of the fact that $`E_1(z;a)=a^z\zeta _R\left(2z\right)`$. Notice that all terms on the R.H.S. of Eq.(10) are zero when $`s0`$ except the last one, that is, $`\zeta \left(0\right)0`$ since $`\zeta _R\left(0\right)=(1/2)\mathrm{ln}(2\pi )`$. This will give rise to a scale-dependent term absent in previous calculations related to this problem . By making use of Eq.( 5) we see that the Helmholtz free energy for the confined electromagnetic field confined within a perfectly conducting rectangular box at finite temperature is $`F(a,b,c,\beta )`$ $`=`$ $`{\displaystyle \frac{abc}{16\pi ^2}}{\displaystyle \underset{l,m,n,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(a^2l^2+b^2m^2+c^2n^2+\frac{p^2\beta ^2}{4}\right)^2}}`$ (11) $`+{\displaystyle \frac{a}{16\pi }}{\displaystyle \underset{l,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(a^2l^2+\frac{p^2\beta ^2}{4}\right)}}+{\displaystyle \frac{b}{16\pi }}{\displaystyle \underset{m,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(b^2m^2+\frac{p^2\beta ^2}{4}\right)}}`$ $`+{\displaystyle \frac{c}{16\pi }}{\displaystyle \underset{n,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(c^2n^2+\frac{p^2\beta ^2}{4}\right)}}+{\displaystyle \frac{1}{\beta }}\mathrm{ln}\left({\displaystyle \frac{\mu \beta }{\sqrt{2\pi }}}\right).`$ The presence of a scale-dependent term is a leftover from the renormalization procedure carried on via the generalized zeta method and introduces an ambiguity in the Helmholtz free energy related to the problem in question. This ambiguity is introduced whenever $`\zeta (s)`$ is not zero at $`s=0`$, see for instance. It was shown by Myers that this ambiguty is a legitimate feature of the generalized zeta function regularization method, whenever the theory is massive or/and interacting and/or some dimensions are compactified. As consequence the usual mode sum method and the generalized zeta function method are not always equivalent. However, the usual mode sum method can be modified so as to include the complexities of a scale dependence . At zero temperature and for the simple geometrical cavity that we are considering here scale-dependent terms are not expected. This is confirmed by the independent calculations due to Lukosz , Ruggiero et al , Ambjorn and Wolfram and the present authors. Here it is clear that the scale dependence is introduced by the generalized zeta function regularization method due to the imposition of periodic conditions on the Euclidean time direction. For our purposes the ambiguity in the free energy can be solved by the additional physical requirement that in unconstrained space, that is, for a very large rectangular box, the only surviving term must be the Stefan-Boltzmann term. It is seen then that $`\mu `$ should be equal to $`\sqrt{2\pi }/\beta `$. ## 3 Temperature inversion symmetry The explicit verification is of temperature inversion symmetry in the case of the system in question is complicated by the fact that in addition to the inverse temperature parameter $`\beta `$ we have to deal with three other length parameters, namely, $`a,b`$ and $`c`$, the measures of the sides of the rectangular cavity. At finite temperature and periodic or antiperiodic conditions along one spatial direction, or Dirichlet, or Neumann plane surfaces located perpendicularly to one spatial direction only two characteristic lengths are involved, a feature that rends this verification easier. Neverheless, by making recourse to a simple trick we shall show that temperature inversion symmetry is also present here. First notice that the denominator of the first term on the R.H.S of Eq.(11) can be rewritten as $`a^2l^2+b^2m^2+c^2n^2=\kappa ^2\left(a^2q^2+b^2r^2+c^2t^2\right)`$, where $`\{q,r,t\}`$ is a sequence of three integers with no common factor and $`\kappa `$ is the common factor of $`\left\{l,m.n\right\}`$. For the sequence $`\{q,r,t\}`$ we define a characteristic length $`d_{\{q,r,t\}}`$ by $$d_{\{q,r,t\}}^2:=a^2q^2+b^2r^2+c^2t^2.$$ (12) Let us also define the dimensionless variable $$\xi _{\{q,r,t\}}:=\frac{2d_{\{q,r,t\}}}{\beta }.$$ (13) The remaining terms on the R.H.S. of Eq.(11) can be treated in a simpler way. For example, for the second term we define $$\xi _a:=\frac{2a}{\beta },$$ (14) Then making the replacement: $`_{l,m,n=\mathrm{}}_{\{q,r,t\}}_{\kappa =\mathrm{}}^+\mathrm{}`$, the free energy density corresponding to Eq.(11) can be written as $`{\displaystyle \frac{F(a,b,c,\beta )}{abc}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle \underset{\{q,r,t\}}{}}{\displaystyle \underset{\kappa ,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{d_{\{q,r,t\}}^4}}{\displaystyle \frac{\xi _{\{q,r,t\}}^4}{\left(\kappa ^2\xi _{\{q,r,t\}}^2+p^2\right)^2}}`$ (15) $`+{\displaystyle \frac{1}{16\pi a^2bc}}{\displaystyle \underset{l,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\xi _a^2}{\left(l_a^2\xi _a^2+p^2\right)}}`$ $`+{\displaystyle \frac{1}{16\pi ab^2c}}{\displaystyle \underset{m,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\xi _b^2}{\left(m_b^2\xi _b^2+p^2\right)}}`$ $`+{\displaystyle \frac{1}{16\pi abc^2}}{\displaystyle \underset{n,p=\mathrm{}}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\xi _c^2}{\left(n^2\xi _c^2+p^2\right)}}`$ The zero temperature limit of Eq.(15) is obtained by reconizing that when $`\beta \mathrm{}`$ the dimensionless parameters $`\xi _{\{q,r,t\}},\xi _a,\xi _b,\xi _c0`$ and the surviving terms in Eq.(15) correspond to $`p=0`$, as the reader can easily verify. Going back to our initial set of indexes $`l,m,n`$ we obtain $$E_0==\frac{abc}{16\pi ^2}\underset{l,m,n=\mathrm{}}{\overset{+\mathrm{}}{}}\frac{1}{\left[a^2l^2+b^2m^2+c^2n^2\right]^2}+\frac{\pi }{48}(\frac{1}{a}+\frac{1}{b}+\frac{1}{c}),$$ (16) which is in perfect agreement with the result obtained by Lukosz and also by Ruggiero et al.. In order to introduce temperature inversion symmetry for this problem let us now define the dimensionless functions $$_{\{q,r,t\}}\left(\xi _{\{q,r,t\}}\right):=\frac{1}{16\pi ^2}\underset{\kappa ,p=\mathrm{}}{\overset{\mathrm{}}{}}\frac{𝒯_{\{q,r,t\}}^4}{\left(\kappa ^2\xi _{\{q,r,t\}}^2+p^2\right)^2},$$ (17) and $$_a\left(\xi _a\right):=\frac{1}{16\pi }\underset{l,p=\mathrm{}}{\overset{\mathrm{}}{}}\frac{\xi _a^2}{\left(l_a^2\xi _a^2+p^2\right)},$$ (18) with similar definitions for $`_b\left(\xi _b\right)`$ and $`_c\left(\xi _c\right)`$. Therefore, in terms of these functions we can write the free energy density as $$\frac{F}{V}=\underset{\{q,r,t\}}{}\frac{_{\{q,r,t\}}\left(\xi _{\{q,r,t\}}\right)}{d_{\{q,r,t\}}^4}+\frac{_a\left(\xi _a\right)}{aV}+\frac{_b\left(𝒯_b\right)}{bV}+\frac{_c\left(\xi _c\right)}{cV},$$ (19) where $`V=abc`$. It is not hard to see that the functions $`_{\{q,r,t\}}\left(\xi _{\{q,r,t\}}\right)`$ exhibit the following property $$\xi _{\{q,r,t\}}^4_{\{q,r,t\}}\left(\frac{1}{\xi _{\{q,r,t\}}}\right)=_{\{q,r,t\}}\left(\xi _{\{q,r,t\}}\right).$$ (20) In the same way $$\xi _a^2f_a\left(\frac{1}{\xi _a}\right)=f_a\left(\xi _a\right),$$ (21) Equations (20) and (21) plus two similar equations for the functions $`_b\left(\xi _b\right)`$ and $`_c\left(\xi _c\right)`$ describe temperature inversion symmetry for the case in question, that is, all terms in Eq.(19) can be inverted by making use of these formulae. In the very high temperature limit we expect the leading contribution to be the Stefan-Boltzmann term, $`\pi ^2/\left(45\beta ^4\right)`$. Now we argue that our transformations are applied in order to generate from the unique Stefan-Boltzmann density term, an infinite number of terms which must be added at zero temperature, and inversely, each term at zero temperature goes to the unique Stefan-Boltzmann term. In this form, that is, as far as the energy density is concerned, we can apply the temperature inversion symmetry transformations to several Casimir systems, including of course the ones previously analyzed in the literature. The specific form of the transformations depends on the particular system at hand, but the procedure is the same. Now, we can easily check that $`_{\{q,r,t\}}\left(0\right)=\pi ^2/720`$, and that $`_a\left(\xi _a\right)=_b\left(\xi _b\right)=_c\left(\xi _c\right)=\pi /48`$. Then, making use of Eqs.(20) and (21) we write $$\xi _{\{q,r,t\}}^4_{\{q,r,t\}}\left(\mathrm{}\right)=\frac{\pi ^2}{720},$$ (22) and $$\xi _a^2_a\left(\mathrm{}\right)=\xi _b^2_b\left(\mathrm{}\right)=\xi _c^2_c\left(\mathrm{}\right)=\frac{\pi }{48}.$$ (23) Taking these results into Eq.(19) we obtain $$F\left(a,b,c,\beta 0\right)\frac{abc\pi ^2}{45\beta ^4}+\frac{\pi }{12\beta ^2}\left(a+b+c\right),$$ (24) which is in agreement with with respect to the leading terms in the high temperature approximation. As expected, the Stefan-Boltzmann term is the leading term in this limit. It is convenient to rewrite Eq.(5), or its equivalent Eq.(19), with the zero temperature terms and the Stefan-Boltzmann term separated from the non-trivial temperature corrections, $`{\displaystyle \frac{F}{V}}`$ $`=`$ $`{\displaystyle \frac{\pi ^2}{45\beta ^4}}+{\displaystyle \frac{\pi }{12\beta ^2}}\left({\displaystyle \frac{1}{ab}}+{\displaystyle \frac{1}{ac}}+{\displaystyle \frac{1}{bc}}\right)`$ (25) $`{\displaystyle \frac{1}{16\pi ^2}}{\displaystyle \underset{l,m,n=\mathrm{}}{\overset{+\mathrm{}}{}}}{\displaystyle \frac{1}{\left[a^2l^2+b^2m^2+c^2n^2\right]^2}}+{\displaystyle \frac{\pi }{48abc}}\left({\displaystyle \frac{1}{a}}+{\displaystyle \frac{1}{b}}+{\displaystyle \frac{1}{c}}\right)`$ $`{\displaystyle \frac{1}{4\pi ^2}}{\displaystyle \underset{\{q,r,t\}}{}}{\displaystyle \underset{\kappa ,p=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(\kappa ^2d_{\{q,r,t\}}^2+\beta ^2p^2\right)^2}}+{\displaystyle \frac{1}{\pi bc}}{\displaystyle \underset{l,p=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(4a^2l^2+\beta ^2p^2\right)}}`$ $`+{\displaystyle \frac{1}{\pi ac}}{\displaystyle \underset{l,p=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(4b^2l^2+\beta ^2p^2\right)}}+{\displaystyle \frac{1}{\pi ab}}{\displaystyle \underset{l,p=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\left(4c^2l^2+\beta ^2p^2\right)}}.`$ The last two lines of Eq.(25) represent the non-trivial temperature-dependent corrections. The first term in the third line is equivalent to a set of conducting plates with each pair of plates characterized by a distance $`d_{\{q,r,t\}}`$, which means that we can take advantage of results already established in the literature regarding high and low temperature approximations for a pair of conducting plates. ## 4 Conclusions In this letter we have employed generalized zeta function techniques in its global version to derive the finite temperature vacuum energy associated with an electromagnetic field confined within a perfectly conducting rectangular box. We have shown that Helmholtz free energy for the problem in question is scale dependent, a feature that was not present in previous calculations concerning this problem at zero temperature , and finite temperature . The scale dependence is a remainder that tell us that the theory was renormalized by the way of the generalized zeta function method. We have also shown that the concept of temperature inversion symmetry can be extended so as to include the elctromagnetic field confined by perfectly conducting rectangular cavity. The main point here is the recognition that the concept of temperature inversion symmetry can be extended in the sense that it must be applied separately to the different terms comprising the non-trivial part of the free energy density, with each term behaving differently under this symmetry. This is the content of Eqs. (20) and (21). The most important feature of temperature inversion symmetry is that it allows us to establish a relationship between the zero and low-temperature sector of the Helmholtz free energy and the high-temperature one. This can be very useful since in general it is easier to obtain low-temperature expansions. In particular, it is possible to relate the zero temperature Casimir effect to the Stefan-Boltzmann term in a straightforward way. ## Appendix Multidimensional homogeneous Epstein function (see also Elizalde et al. for a modern presentation) are defined by $$E_N(z;a_1,a_2,\mathrm{}a_N)=\underset{n_1,n_2,..n_N=1}{\overset{\mathrm{}}{}}\left(a_1n_1^2+a_2n_2^2+\mathrm{}+a_Nn_N^2\right)^z,$$ $`(A.1)`$ for $`\mathrm{}z>N/2`$ and $`a_1,a_2,\mathrm{}a_N>\mathrm{\hspace{0.17em}0}`$. Generalized zeta functions. $`A_N(z;a_1,a_2,\mathrm{}a_N)`$ are defined by , $$A_N(z;a_1,a_2,\mathrm{}a_N)=\underset{n_1,n_2,..n_N=\mathrm{}}{\overset{+\mathrm{}}{}}\left(a_1n_1^2+a_2n_2^2+\mathrm{}+a_Nn_N^2\right)^z,$$ $`(A.2)`$ for $`\mathrm{}z>N/2`$ and $`a_1,a_2,\mathrm{}a_N>\mathrm{\hspace{0.17em}0}`$. The prime means that the term corresponding to $`n_1=n_2=\mathrm{}=n_N=0`$ must be excluded from the summation. A useful reflection formula reads envolving these functions is $$A_N(z;a_1,a_2,\mathrm{}a_N)=\frac{\pi ^{\frac{N}{2}+2z}}{\sqrt{a_1\mathrm{}a_N}}\frac{\mathrm{\Gamma }\left(\frac{N}{2}z\right)}{\mathrm{\Gamma }\left(z\right)}A_N(\frac{N}{2}z;1/a_1,1/a_2,\mathrm{..1}/.a_N).$$ $`(A.3)`$ Relationships between generalized zeta functions and Epstein functions can be derived from the very definition given by ($`(A.2)`$). For example, $$A_2(z;a_1,a_2)=2^2E_2(z;a_1,a_2)+2E_1(z;a_1)+2E_1(z;a_2)$$ $`(A.4)`$ and $`A_4(z;a_1,a_2,a_3,a_4)=2^4E_4(z;a_1,a_2,a_3,a_4)+2^3E_3(z;a_1,a_2,a_3)`$ $`+2^3E_3(z;a_1,a_2,a_4)+2^3E_3(z;a_1,a_3,a_4)+2^2E_2(z;a_1,a_2)+2^2E_2(z;a_1,a_3)`$ $`+2^3E_3(z;a_2,a_3,a_4)+2^2E_2(z;a_2,a_3)+2^2E_2(z;a_1,a_4)+2^2E_2(z;a_2,a_4)`$ $`+2^2E_2(z;a_3,a_4)+2E_1(z;a_1)+2E1(z;a_2)+2E1(z;a_3)+2E_1(z;a_4),(A.5)`$ of which use was made in this letter. ### Acknowledgments The authors wish to acknowledge useful converstions with Dr. A. A. Actor.
warning/0003/nucl-th0003002.html
ar5iv
text
# Equilibration within a semiclassical off-shell transport approach11footnote 1supported by GSI Darmstadt ## 1 Introduction The dynamical description of strongly interacting systems out of equilibrium nowadays is dominantly based on transport theories and efficient numerical recipies have been set up for the solution of the coupled channel transport equations (and Refs. therein). These transport approaches have been derived either from the Kadanoff-Baym equations in Refs. or from the hierarchy of connected equal-time Green functions in Refs. by applying a Wigner transformation and restricting to first order in the derivatives of the phase-space variables ($`X,P`$). Whereas theoretical formulations of off-shell quantum transport have been limited to the formal level for a couple of years only recently a tractable semiclassic form has been derived for testparticles in the eight dimensional phase space of a particle . Whereas in Refs. we have investigated the off-shell transport approach with respect to nucleus-nucleus collisions at GANIL, SIS and AGS energies, we here concentrate on equilibration phenomena relative to the on-shell dynamics by imposing periodic boundary conditions for the system confined to a box of size $`V=L^3`$, where $`L`$ denotes the length of the cubic box. We, furthermore, compare the equilibrium nucleon and $`\mathrm{\Delta }`$ distribution functions ($`t\mathrm{}`$) to the statistical model (SM) employing the same spectral functions as in the transport approach. For related studies at higher bombarding energies or energy densities within on-shell transport approaches we refer the reader to Refs. . ## 2 Extended semiclassical transport equations We briefly recall the basic equations for Green functions and particle self energies that are exploited in the derivation of off-shell transport equations in the semiclassical limit. The general starting point for the derivation of a transport equation for particles with a finite and dynamical width are the Dyson-Schwinger equations for the retarded and advanced Green functions $`S^{ret}`$, $`S^{adv}`$ and for the non-ordered Green functions $`S^<`$ and $`S^>`$ . In the case of scalar bosons – which is considered in the following for simplicity – these Green functions are defined by $`iS_{xy}^<`$ $`:=`$ $`<\mathrm{\Phi }^{}(y)\mathrm{\Phi }(x)>,iS_{xy}^>:=<\mathrm{\Phi }(x)\mathrm{\Phi }^{}(y)>,`$ $`iS_{xy}^{ret}`$ $`:=`$ $`\mathrm{\Theta }(x_0y_0)<[\mathrm{\Phi }(x),\mathrm{\Phi }^{}(y)]>,`$ $`iS_{xy}^{adv}`$ $`:=`$ $`\mathrm{\Theta }(y_0x_0)<[\mathrm{\Phi }(x),\mathrm{\Phi }^{}(y)]>.`$ (1) They depend on the space-time coordinates $`x,y`$ as indicated by the indices $`_{xy}`$. The Green functions are determined via Dyson-Schwinger equations by the retarded and advanced self energies $`\mathrm{\Sigma }^{ret},\mathrm{\Sigma }^{adv}`$ and the collisional self energy $`\mathrm{\Sigma }^<`$: $`\widehat{S}_{0x}^1S_{xy}^{ret}=\delta _{xy}+\mathrm{\Sigma }_{xz}^{ret}S_{zy}^{ret},`$ (2) $`\widehat{S}_{0x}^1S_{xy}^{adv}=\delta _{xy}+\mathrm{\Sigma }_{xz}^{adv}S_{zy}^{adv},`$ (3) $`\widehat{S}_{0x}^1S_{xy}^<=\mathrm{\Sigma }_{xz}^{ret}S_{zy}^<+\mathrm{\Sigma }_{xz}^<S_{zy}^{adv},`$ (4) where Eq. (4) is the well-known Kadanoff-Baym equation. Here $`\widehat{S}_{0x}^1`$ denotes the (negative) Klein-Gordon differential operator which is given for bosonic field quanta of (bare) mass $`M_0`$ by $`\widehat{S}_{0x}^1=(_x^\mu _\mu ^x+M_0^2)`$; $`\delta _{xy}`$ represents the four-dimensional $`\delta `$-distribution $`\delta _{xy}\delta ^{(4)}(xy)`$ and the symbol $``$ indicates an integration (from $`\mathrm{}`$ to $`\mathrm{}`$) as well as a summation over all discrete intermediate variables (cf. ). ### 2.1 The semiclassical limit For the derivation of a semiclassical transport equation one now changes from a pure space-time formulation into the Wigner-representation. The theory is then formulated in terms of the center-of-mass variable $`X=(x+y)/2`$ and the momentum $`P`$, which is introduced by Fourier-transformation with respect to the relative space-time coordinate $`(xy)`$. In any semiclassical transport theory one, furthermore, keeps only contributions up to the first order in the space-time gradients. After carrying-out these two steps the Dyson-Schwinger equations (2)-(4) become $`\left[P^2M_0^2+iP^\mu _\mu ^X\right]S_{XP}^{ret}=\mathrm{\hspace{0.33em}1}+(\mathrm{\hspace{0.17em}1}i\mathrm{})\{\mathrm{\Sigma }_{XP}^{ret}\}\{S_{XP}^{ret}\},`$ (5) $`\left[P^2M_0^2+iP^\mu _\mu ^X\right]S_{XP}^{adv}=\mathrm{\hspace{0.33em}1}+(\mathrm{\hspace{0.17em}1}i\mathrm{})\{\mathrm{\Sigma }_{XP}^{adv}\}\{S_{XP}^{adv}\},`$ (6) $`\left[P^2M_0^2+iP^\mu _\mu ^X\right]S_{XP}^<=(\mathrm{\hspace{0.17em}1}i\mathrm{})\left[\{\mathrm{\Sigma }_{XP}^{ret}\}\{S_{XP}^<\}+\{\mathrm{\Sigma }_{XP}^<\}\{S_{XP}^{adv}\}\right],`$ (7) where the operator $`\mathrm{}`$ is defined as $`\mathrm{}\{F_1\}\{F_2\}:={\displaystyle \frac{1}{2}}\left({\displaystyle \frac{F_1}{X^\mu }}{\displaystyle \frac{F_2}{P_\mu }}{\displaystyle \frac{F_1}{P_\mu }}{\displaystyle \frac{F_2}{X^\mu }}\right).`$ (8) It is a four-dimensional generalization of the well-known Poisson-bracket. Starting from (5) and (6) one obtains algebraic relations between the real and the imaginary part of the retarded Green functions. On the other hand eq. (7) leads to a ’transport equation’ for the Green function $`S^<`$ . To this aim one separates all retarded and advanced quantities – Green functions and self energies – into real and imaginary parts, $`S_{XP}^{ret,adv}=ReS_{XP}^{ret}{\displaystyle \frac{i}{2}}A_{XP},\mathrm{\Sigma }_{XP}^{ret,adv}=Re\mathrm{\Sigma }_{XP}^{ret}{\displaystyle \frac{i}{2}}\mathrm{\Gamma }_{XP}.`$ (9) The imaginary part of the retarded propagator is given (up to a factor 2) by the normalized spectral function $`A_{XP}=i\left[S_{XP}^{ret}S_{XP}^{adv}\right]=2ImS_{XP}^{ret},{\displaystyle \frac{dP_0^2}{4\pi }A_{XP}}=\mathrm{\hspace{0.33em}1},`$ (10) while the imaginary part of the self energy corresponds to half the particle width $`\mathrm{\Gamma }_{XP}`$. By separating the complex equations (5) and (6) into their real and imaginary contributions we obtain an algebraic equation between the real and the imaginary part of $`S^{ret}`$, $`ReS_{XP}^{ret}={\displaystyle \frac{P^2M_0^2Re\mathrm{\Sigma }_{XP}^{ret}}{\mathrm{\Gamma }_{XP}}}A_{XP}.`$ (11) In addition we gain an algebraic solution for the spectral function as $`A_{XP}={\displaystyle \frac{\mathrm{\Gamma }_{XP}}{(P^2M_0^2Re\mathrm{\Sigma }_{XP}^{ret})^2+\mathrm{\Gamma }_{XP}^2/4}},`$ (12) while the real part of the retarded propagator is given by $`ReS_{XP}^{ret}={\displaystyle \frac{P^2M_0^2Re\mathrm{\Sigma }_{XP}^{ret}}{(P^2M_0^2Re\mathrm{\Sigma }_{XP}^{ret})^2+\mathrm{\Gamma }_{XP}^2/4}}.`$ (13) Furthermore, the (Wigner-transformed) Kadanoff-Baym equation (7) allows for the construction of a transport equation for the Green function $`S^<`$. When separating the real and the imaginary contribution of this equation we find i) a generalized transport equation, $`\mathrm{}\{P^2`$ $``$ $`M_0^2Re\mathrm{\Sigma }_{XP}^{ret}\}\{S_{XP}^<\}\mathrm{}\{\mathrm{\Sigma }_{XP}^<\}\{ReS_{XP}^{ret}\}`$ (14) $`=`$ $`{\displaystyle \frac{i}{2}}\left[\mathrm{\Sigma }_{XP}^>S_{XP}^<\mathrm{\Sigma }_{XP}^<S_{XP}^>\right],`$ and ii) a generalized mass-shell constraint $`[P^2`$ $``$ $`M_0^2Re\mathrm{\Sigma }_{XP}^{ret}]S^<_{XP}\mathrm{\Sigma }^<_{XP}ReS^{ret}_{XP}`$ (15) $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{}\{\mathrm{\Sigma }_{XP}^<\}\{A_{XP}\}{\displaystyle \frac{1}{2}}\mathrm{}\{\mathrm{\Gamma }_{XP}\}\{S_{XP}^<\}.`$ Besides the drift term (i.e. $`\mathrm{}\{P^2M_0^2\}\{S^<\}=P^\mu _\mu ^XS^<)`$ and the Vlasov term (i.e. $`\mathrm{}\{Re\mathrm{\Sigma }^{ret}\}\{S^<\}`$) a third contribution appears on the l.h.s. of (14) (i.e. $`\mathrm{}\{\mathrm{\Sigma }^<\}\{ReS^{ret}\}`$), which vanishes in the quasiparticle limit and incorporates – as shown in – the off-shell behaviour in the particle propagation which has been neglected so far in transport studies<sup>2</sup><sup>2</sup>2This also holds true for the recent numerical off-shell simulations in Refs. . The r.h.s. of (14) consists of a collision term with its characteristic gain ($`\mathrm{\Sigma }^<S^>`$) and loss ($`\mathrm{\Sigma }^>S^<`$) structure, where scattering and decay processes of particles are described. Within the specific term ($`\mathrm{}\{\mathrm{\Sigma }^<\}\{ReS^{ret}\}`$) a further modification is necessary. According to Botermans and Malfliet the collisional self energy $`\mathrm{\Sigma }^<`$ should be replaced by $`S^<\mathrm{\Gamma }/A`$ to gain a consistent first order gradient expansion scheme. The replacement is allowed since the difference between these two expressions $`(\mathrm{\Sigma }^<S^<\mathrm{\Gamma }/A)`$ can be shown to be of first order in the space-time gradients itself . Furthermore, this substitution is required to get rid of the inequivalence between the general transport equation (14) and the general mass shell constraint (15) . Finally, the general transport equation (in first order gradient expansion) reads $`A_{XP}\mathrm{\Gamma }_{XP}`$ $`\left[\mathrm{}\{P^2M_0^2Re\mathrm{\Sigma }_{XP}^{ret}\}\{S_{XP}^<\}{\displaystyle \frac{1}{\mathrm{\Gamma }_{XP}}}\mathrm{}\{\mathrm{\Gamma }_{XP}\}\{(P^2M_0^2Re\mathrm{\Sigma }_{XP}^{ret})S_{XP}^<\}\right]`$ (16) $`=i\left[\mathrm{\Sigma }_{XP}^>S_{XP}^<\mathrm{\Sigma }_{XP}^<S_{XP}^>\right].`$ Its formal structure is fixed by the approximations applied, however, its physical contents is fully determined by the different self energies, i.e. $`Re\mathrm{\Sigma }_{XP}^{ret},\mathrm{\Gamma }_{XP},\mathrm{\Sigma }_{XP}^<`$ that have to be specified in addition. In order to obtain an approximate solution to the transport equation (16) a testparticle ansatz is used for the Green function $`S^<`$, more specifically for the real and positive semidefinite quantity $`F_{XP}=A_{XP}N_{XP}=iS_{XP}^<`$, $`F_{XP}{\displaystyle \underset{i=1}{\overset{N}{}}}\delta ^{(3)}(\stackrel{}{X}\stackrel{}{X}_i(t))\delta ^{(3)}(\stackrel{}{P}\stackrel{}{P}_i(t))\delta (P_0ϵ_i(t)).`$ (17) In the most general case (where the self energies depend on four-momentum $`P`$, time $`t=X_0`$ and the spatial coordinates $`\stackrel{}{X}`$) the equations of motion for the testparticles read $`{\displaystyle \frac{d\stackrel{}{X}_i}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{1C_{(i)}}}{\displaystyle \frac{1}{2ϵ_i}}\left[\mathrm{\hspace{0.17em}2}\stackrel{}{P}_i+\stackrel{}{}_{P_i}Re\mathrm{\Sigma }_{(i)}^{ret}+{\displaystyle \frac{ϵ_i^2\stackrel{}{P}_i^2M_0^2Re\mathrm{\Sigma }_{(i)}^{ret}}{\mathrm{\Gamma }_{(i)}}}\stackrel{}{}_{P_i}\mathrm{\Gamma }_{(i)}\right],`$ (18) $`{\displaystyle \frac{d\stackrel{}{P}_i}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{1C_{(i)}}}{\displaystyle \frac{1}{2ϵ_i}}\left[\stackrel{}{}_{X_i}Re\mathrm{\Sigma }_i^{ret}+{\displaystyle \frac{ϵ_i^2\stackrel{}{P}_i^2M_0^2Re\mathrm{\Sigma }_{(i)}^{ret}}{\mathrm{\Gamma }_{(i)}}}\stackrel{}{}_{X_i}\mathrm{\Gamma }_{(i)}\right],`$ (19) $`{\displaystyle \frac{dϵ_i}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{1C_{(i)}}}{\displaystyle \frac{1}{2ϵ_i}}\left[{\displaystyle \frac{Re\mathrm{\Sigma }_{(i)}^{ret}}{t}}+{\displaystyle \frac{ϵ_i^2\stackrel{}{P}_i^2M_0^2Re\mathrm{\Sigma }_{(i)}^{ret}}{\mathrm{\Gamma }_{(i)}}}{\displaystyle \frac{\mathrm{\Gamma }_{(i)}}{t}}\right],`$ (20) where the notation $`F_{(i)}`$ implies that the function is taken at the coordinates of the testparticle, i.e. $`F_{(i)}F(t,\stackrel{}{X}_i(t),\stackrel{}{P}_i(t),ϵ_i(t))`$. In (18-20) a common multiplication factor $`(1C_{(i)})^1`$ appears, which contains the energy derivatives of the retarded self energy $`C_{(i)}={\displaystyle \frac{1}{2ϵ_i}}\left[{\displaystyle \frac{}{ϵ_i}}Re\mathrm{\Sigma }_{(i)}^{ret}+{\displaystyle \frac{ϵ_i^2\stackrel{}{P}_i^2M_0^2Re\mathrm{\Sigma }_{(i)}^{ret}}{\mathrm{\Gamma }_{(i)}}}{\displaystyle \frac{}{ϵ_i}}\mathrm{\Gamma }_{(i)}\right],`$ (21) which yields a shift of the system time $`t`$ to the ’eigentime’ of particle $`i`$ defined by $`\stackrel{~}{t}_i=t/(1C_{(i)})`$. The derivatives with respect to the ’eigentime’, i.e. $`d\stackrel{}{X}_i/d\stackrel{~}{t}_i`$, $`d\stackrel{}{P}_i/d\stackrel{~}{t}_i`$ and $`dϵ_i/d\stackrel{~}{t}_i`$ then emerge without this renormalization factor for each testparticle $`i`$ when neglecting the explicit time dependence of $`C_{(i)}`$ in line with the semiclassical approximation scheme. The role and the importance of this correction factors have been studied in Ref. for a four-momentum-dependent ’trial’ potential and we refer the reader to the latter analysis for more details. Following Ref. we take $`M^2=P^2Re\mathrm{\Sigma }^{ret}`$ as an independent variable instead of the energy $`P_0`$. Eq. (20) then turns to $`{\displaystyle \frac{dM_i^2}{dt}}={\displaystyle \frac{M_i^2M_0^2}{\mathrm{\Gamma }_{(i)}}}{\displaystyle \frac{d\mathrm{\Gamma }_{(i)}}{dt}}`$ (22) for the time evolution of the test-particle $`i`$ in the invariant mass squared as derived in Refs. . We mention that corresponding equations of motion have been derived by Leupold in the nonrelativistic limit . ### 2.2 Generalized collision terms for bosons and fermions The collision term of the Kadanoff-Baym equation can only be worked out in more detail by giving explicit approximations for $`\mathrm{\Sigma }^<`$ and $`\mathrm{\Sigma }^>`$. We recall the formulation and result from Ref. that is based on Dirac-Brueckner theory, i.e. $$I_{coll}(X,\stackrel{}{P},M^2)=Tr_2Tr_3Tr_4A(X,\stackrel{}{P},M^2)A(X,\stackrel{}{P}_2,M_2^2)A(X,\stackrel{}{P}_3,M_3^2)A(X,\stackrel{}{P}_4,M_4^2)$$ $$|T((\stackrel{}{P},M^2)+(\stackrel{}{P}_2,M_2^2)(\stackrel{}{P}_3,M_3^2)+(\stackrel{}{P}_4,M_4^2))|_{𝒜,𝒮}^2\delta ^{(4)}(P+P_2P_3P_4)$$ $$[N_{X\stackrel{}{P}_3M_3^2}N_{X\stackrel{}{P}_4M_4^2}\overline{f}_{X\stackrel{}{P}M^2}\overline{f}_{X\stackrel{}{P}_2M_2^2}N_{X\stackrel{}{P}M^2}N_{X\stackrel{}{P}_2M_2^2}\overline{f}_{X\stackrel{}{P}_3M_3^2}\overline{f}_{X\stackrel{}{P}_4M_4^2}]$$ (23) with $$\overline{f}_{X\stackrel{}{P}M^2}=1+\eta N_{X\stackrel{}{P}M^2}$$ (24) and $`\eta =\pm 1`$ for bosons/fermions, respectively. The indices $`𝒜,𝒮`$ stand for the antisymmetric/symmetric matrix element of the in-medium off-shell scattering amplitude $`T`$ in case of fermions/bosons. In Eq. (23) the trace over particles 2,3,4 reads explicitly for fermions $$Tr_2=\underset{\sigma _2,\tau _2}{}\frac{1}{(2\pi )^4}d^3P_2\frac{dM_2^2}{2\sqrt{\stackrel{}{P}_2^2+M_2^2}},$$ (25) where $`\sigma _2,\tau _2`$ denote the spin and isospin of particle 2. In case of bosons we have instead $$Tr_2=\underset{\sigma _2,\tau _2}{}\frac{1}{(2\pi )^4}d^3P_2\frac{dP_{0,2}^2}{2},$$ (26) since here the spectral function $`A_B`$ is normalized as $$\frac{dP_0^2}{4\pi }A_B(X,P)=1$$ (27) whereas for fermions we have $$\frac{dP_0}{2\pi }A_F(X,P)=1.$$ (28) We mention that the spectral function $`A_F`$ in case of fermions in (23) is obtained by considering only particles of positive energy and assuming the spectral function to be identical for spin ’up’ and ’down’ states (cf. Ref. ). Neglecting the ’gain-term’ in Eq. (23) one recognizes that the collisional width $`\mathrm{\Gamma }_{coll}`$ of the particle in the rest frame is given by $$\mathrm{\Gamma }_{coll}(X,\stackrel{}{P},M^2)=Tr_2Tr_3Tr_4|T((\stackrel{}{P},M^2)+(\stackrel{}{P}_2,M_2^2)(\stackrel{}{P}_3,M_3^2)+(\stackrel{}{P}_4,M_4^2))|_{𝒜,𝒮}^2$$ (29) $$A(X,\stackrel{}{P}_2,M_2^2)A(X,\stackrel{}{P}_3,M_3^2)A(X,\stackrel{}{P}_4,M_4^2)\delta ^4(P+P_2P_3P_4)N_{X\stackrel{}{P}_2M_2^2}\overline{f}_{X\stackrel{}{P}_3M_3^2}\overline{f}_{X\stackrel{}{P}_4M_4^2},$$ where as in Eq. (23) local off-shell transition amplitudes enter for the transitions $`P+P_2P_3+P_4`$. We note that the extension of Eq. (23) to inelastic scattering processes (e.g. $`NNN\mathrm{\Delta }`$) or ($`\pi N\mathrm{\Delta }`$ etc.) is straightforward when exchanging the elastic transition amplitude $`T`$ by the corresponding inelastic one and taking care of Pauli-blocking or Bose-enhancement for the particles in the final state. The relation of the quantity $`\mathrm{\Gamma }_{XP}`$ to the collisional width $`\mathrm{\Gamma }_{coll}`$ is given by $`\mathrm{\Gamma }_{XP}=2P_0(\mathrm{\Gamma }_{decay}(XP)+\mathrm{\Gamma }_{coll}(XP))`$, where the particle decay width $`\mathrm{\Gamma }_{decay}`$ in the medium might also be different compared to the vacuum. Thus the transport approach determines the particle spectral function dynamically via (29) – with respect to the collisional width $`\mathrm{\Gamma }_{coll}`$ – for all hadrons if the in-medium transition amplitudes $`T`$ are known in their full off-shell dependence. Since this information is not available for configurations of hot and dense matter, a couple of assumptions and numerical approximation schemes have to be invoked in actual applications so far. As in Refs. the following dynamical calculations are based on the conventional HSD transport approach – in which $`Re\mathrm{\Sigma }_{XP}^{ret}`$ is specified for the hadrons – however, the equations of motion for the testparticles are extended to (18,19,22). For further details on the elastic and inelastic transition rates we refer the reader to Ref. . ### 2.3 Comment on particle number conservation In previous derivations of the off-shell transport equations one has started from a formulation of the non-equilibrium theory in space-time representation $`(x,x^{})`$ and then changed into a phase-space representation via Fourier transformation with respect to $`(xx^{})`$ . The semiclassical limit then has been introduced by assuming gradients in $`X=(x+x^{})/2`$ to be small for $`Re\mathrm{\Sigma }_{XP}^{ret}`$ and $`S_{XP}^<`$. Here we argue that for reasons of symmetry in phase-space also the four-momentum derivatives in $`P`$ have to be small to achieve a proper semiclassical limit as inherent in the formulation of the transport equation (16) in terms of the generalized Poisson-bracket (8). We briefly demonstrate in the following lines that instead of a coordinate-space representation one may formulate the theory equivalently in momentum-space and then transform to phase space: The momentum-dependent (two-point) functions are given as $`F_{p_1p_1^{}}={\displaystyle d^4x_1d^4x_1^{}e^{i(p_1x_1p_1^{}x_1^{})}F_{x_1x_1^{}}}.`$ (30) The evolution of the retarded and advanced Green functions $`S^{ret},S^{adv}`$ and the Green function $`S^<`$ turns to $`\widehat{S}_{0p_1}^1S_{p_1p_1^{}}^{ret,adv}`$ $`=`$ $`(2\pi )^4\delta ^{(4)}(p_1p_1^{})+{\displaystyle \frac{d^4p_2}{(2\pi )^4}\mathrm{\Sigma }_{p_1p_2}^{ret,adv}S_{p_2p_1^{}}^{ret,adv}},`$ (31) $`\widehat{S}_{0p_1}^1S_{p_1p_1^{}}^<`$ $`=`$ $`{\displaystyle \frac{d^4p_2}{(2\pi )^4}\left[\mathrm{\Sigma }_{p_1p_2}^{ret}S_{p_2p_1^{}}^<+\mathrm{\Sigma }_{p_1p_2}^<S_{p_2p_1^{}}^{adv}\right]},`$ (32) with the ’kinetic’ operator in momentum-space $`\widehat{S}_{0p_1}^1=(p_1^2M_0^2)`$ in the case of relativistic scalar bosons with (bare) mass $`M_0`$. Obviously, the equations in momentum-space are formally equivalent to the equations in coordinate-space, i.e. the convolution integrals in the coordinate $`x_2`$ are replaced by convolution integrals in the momentum $`p_2`$. When transforming from momentum- to phase-space via a Fourier-transformation with respect to the four-momentum coordinate $`(p_1p_1^{})`$ $`F_{XP}={\displaystyle \frac{d^4(p_1p_1^{})}{(2\pi )^4}e^{iX(p_1p_1^{})}F_{p_1p_1^{}}}`$ (33) (with $`P=(p_1+p_1^{})/2`$) we gain again the familiar equations in Wigner-representation: $`\widehat{S}_{0XP}^1S_{XP}^{ret,adv}`$ $`=`$ $`\mathrm{\hspace{0.33em}1}+e^i\mathrm{}\mathrm{\Sigma }_{XP}^{ret,adv}S_{XP}^{ret,adv}`$ (34) $`\widehat{S}_{0XP}^1S_{XP}^<`$ $`=`$ $`e^i\mathrm{}\left[\mathrm{\Sigma }_{XP}^{ret}S_{XP}^<+\mathrm{\Sigma }_{XP}^<S_{XP}^{adv}\right],`$ (35) with the ’kinetic’ operator in phase-space $`\widehat{S}_{0XP}^1=(P^2+iP^\mu _\mu ^X\frac{1}{4}_\mu ^X_X^\mu M_0^2)`$. Here we have used the following relation for the convolution integrals in momentum-space $`{\displaystyle \frac{d^4(p_1p_1^{})}{(2\pi )^4}e^{iX(p_1p_1^{})}\frac{d^4p_2}{(2\pi )^4}F_{1,p_1p_2}F_{2,p_2p_1^{}}}=e^i\mathrm{}F_{1,PX}F_{2,PX},`$ (36) with the Poisson-operator $`\mathrm{}`$ defined in (8). The semiclassical limit – along the conventional line of arguments – now is achieved by assuming gradients of the self energies $`\mathrm{\Sigma }^{ret,adv}`$ in $`P`$ to be small. This assumption is especially well taken for systems with dominantly short range interactions since the different self energies $`\mathrm{\Sigma }^{ret,adv}(p,p^{})`$ become smooth functions in momentum-space such that a restriction to first order gradients in the momentum $`P=(p+p^{})/2`$ can be more easily justified. Note again that the four-dimensional Poisson-bracket (8) is symmetric in the space-time and the four-momentum derivatives. Thus a phase-space gradient expansion requires all gradients to be small contrary to the assumptions made in Ref. . As a consequence mixed gradients of second order as $`^2/(tp_0)`$ have to be neglected in a consistent truncation scheme of first order. As shown by Botermans and Malfliet in the review the particle number conservation then holds strictly. Only when keeping special terms of second order $`(^2/(tp_0)`$ unphysical pecularities may appear, that lead to a violation of particle number conservation in contradiction to the full theory (2)-(4). ## 3 Infinite nuclear matter problems In case of infinite nuclear matter problems the solution of the transport equations simplifies since all spatial gradients (with respect to $`\stackrel{}{X}`$) vanish. This implies that the momentum coordinates of the testparticles $`\stackrel{}{P}_i`$ are constant according to (19) in between collisions. The initial conditions of the problem then are fully specified by an initial bombarding energy per nucleon, that defines the relative shift in momentum of two Fermi spheres with a Fermi momentum $`P_F`$ = 260 MeV/c, the number of nucleons $`N`$ and the total volume of the box $`V=L^3`$. Alternatively, one can also characterize the system by an average density $`\rho =N/V`$ and energy density $`ϵ`$ (cf. Ref. ). We use $`L=10`$ fm and assume an equal number of neutrons and protons. In general (for $`t\mathrm{}`$) the stationary solution of Eq. (23) for a fermion $`h`$ is given by $$F_h(\stackrel{}{X},\stackrel{}{P},M^2)=\frac{A_h(\stackrel{}{X},\stackrel{}{P},M^2)}{\mathrm{exp}((E\mu _h)/T)\eta }$$ (37) with $`E=\sqrt{\stackrel{}{P}^2+M^2}`$ and $`\eta =\pm 1`$ for bosons/fermions, respectively, while $`A_h`$ denotes the spectral function (12), $`T`$ the temperature of the system and $`\mu _h`$ the chemical potential for the hadron. This situation corresponds to the grand canonical ensemble of quantum statistical mechanics. In case of infinite nuclear matter problems the dependence on $`\stackrel{}{X}`$ vanishes additionally. ### 3.1 Numerical results As an example for equilibration phenomena we show in Fig. 1 the time evolution of the quadrupole moment (involving all baryons $`B`$) $$Q_2(t)=\underset{B}{}\frac{g_B}{(2\pi )^4}d^3Xd^3P\frac{dM^2}{2\sqrt{\stackrel{}{P}^2+M^2}}(2P_z^2P_x^2P_y^2)F_B(\stackrel{}{X},t,\stackrel{}{P},M^2),$$ (38) for $`\rho =\rho _0`$ and initial bombarding energies of 0.1 A GeV and 1 A GeV, respectively. In (38) the degeneracy factor is $`g_B=4`$ for nucleons and $`g_B`$= 16 for $`\mathrm{\Delta }`$-resonances. The solid lines result from the off-shell calculation while the dashed lines result from the on-shell limit. As can be seen from Fig. 1 the off-shell results are practically identical to the on-shell limits except for the very long time behaviour, where the off-shell limit needs some more time to achieve equilibrium. However, when defining an equilibration time $`\tau `$ by a drop of the observable by the factor $`e^1`$ we find no sizeable effect from the off-shell propagation on $`\tau `$. The equilibrium distributions in the nucleon transverse mass $`M_t=\sqrt{p_t^2+M^2}`$ are shown in Fig. 2 for initial bombarding energies of 0.1 A GeV (upper part) and 1 A GeV (lower part). Again the solid lines denote the off-shell results from the transport calculations while the on-shell spectra are displayed in terms of dashed lines. Both limits (within the statistics) give the same temperature $`T`$ 97 MeV for 1.0 A GeV and $`T`$ 26 MeV for 0.1 A GeV as can be extracted from the high energy tail of the transverse mass spectra. Differences can only be found for $`m_TM_0`$ since the finite width in the nucleon spectral function only can show up in the off-shell case. This component is quite small at 0.1 A GeV since the collisional width of nucleons at density $`\rho _0`$ and temperature $`T`$ 26 MeV is about 8 MeV. Without explicit display we mention that the time evolution of the nucleon spectral function in the invariant mass $`M`$ becomes broad in the initial nonequilibrium phase of the reaction and approaches the equilibrium distribution (37) roughly within the equilibration times from Fig. 1. Since the width of the nucleon spectral function in equilibrium at 0.1 A GeV (GANIL energy) is rather small ($`\mathrm{\Gamma }`$ 8 MeV) we concentrate on the energy of 1 A GeV (SIS energy) in the following, where the nucleon collisional width $`\mathrm{\Gamma }_{coll}`$ at a temperature of 97 MeV is about 40 MeV and roughly 20% of the baryons are excited $`\mathrm{\Delta }`$ resonances<sup>3</sup><sup>3</sup>3Note, that in collisions of finite systems at this bombarding energy the $`\mathrm{\Delta }`$-abundancy is slightly lower due to a rapid expansion of the system and the additional compressional energy stored in the system in the high density phase.. As demonstrated above, for the equilibrated system we can extract a temperature $`T`$ by fitting the particle spectra with the Bolzmann distribution $`{\displaystyle \frac{d^3N_i}{dp^3}}\mathrm{exp}(E_i/T),`$ (39) where $`E_i=\sqrt{p_i^2+m_i^2}`$ is the energy of particle $`i`$. We note that at the temperatures of interest here, the Bose and Fermi distributions are practically identical to a Boltzmann distribution. We find that in equilibrium the spectra of all particles (nucleons, $`\mathrm{\Delta }`$’s and pions) can be characterized by a single temperature $`T`$ (cf. e.g. Ref. ). ### 3.2 Comparison to the statistical model In order to investigate the equilibrium behaviour of hadron matter we compare our transport (box) calculations with a simple Statistical Model (SM) for an Ideal Hadron Gas where the system is described by a grand canonical ensemble of non-interacting fermions and bosons in equilibrium at temperature $`T`$. All baryon and meson species considered in the transport model ($`N,\mathrm{\Delta },\pi `$) also are included in the statistical model. We recall that in the SM particle multiplicities $`n_i`$ and energy densities $`\epsilon _i`$ for particles with spectral functions $`A_i`$ are given by $`n_i={\displaystyle \frac{g_i}{(2\pi \mathrm{})^3}}{\displaystyle \frac{dM}{2\pi }\underset{0}{\overset{\mathrm{}}{}}\frac{A_i(M)/4\pi p^2dp}{\mathrm{exp}\left[(E_iB_i\mu _BS_i\mu _S)/T\right]\eta }},`$ (40) $`\epsilon _i={\displaystyle \frac{g_i}{(2\pi \mathrm{})^3}}{\displaystyle \frac{dM}{2\pi }\underset{0}{\overset{\mathrm{}}{}}\frac{A_i(M)4\pi E_ip^2dp}{\mathrm{exp}\left[(E_iB_i\mu _BS_i\mu _S)/T\right]\eta }},`$ (41) where $`E_i=\sqrt{p^2+M^2}`$ is the energy of particle $`i`$, $`B_i`$ is the baryon charge, $`S_i`$ is the strangeness, and $`g_i`$ is the spin-isospin degeneracy factor, while $`\eta =\pm 1`$ for bosons/fermions, respectively. In Eqs. (40),(41) $`\mu _B`$ and $`\mu _S`$ are the baryon and strangeness chemical potentials. The energy density $`\epsilon `$, baryon density $`\rho `$ and strange particle density of the whole system in equilibrium then is given as $`\epsilon ={\displaystyle \underset{i}{}}\epsilon _i(T,\mu _B,\mu _S)`$ (42) $`\rho ={\displaystyle \underset{i}{}}B_in_i(T,\mu _B,\mu _S)`$ (43) $`\rho _S={\displaystyle \underset{i}{}}S_in_i(T,\mu _B,\mu _S)\mathrm{\hspace{0.17em}0}.`$ (44) As ’input’ for the SM we use the same $`\epsilon ,\rho `$ and $`\rho _S`$ as in the box calculations and we obtain the thermodynamical parameters – $`T,\mu _B,\mu _S`$ – by solving the system of nonlinear equations (42),(43) and (44). We now turn to a comparison of the equilibrium distributions in mass for nucleons and $`\mathrm{\Delta }`$’s, i.e. $`dN_N/dM`$ and $`dN_\mathrm{\Delta }/dM`$, from the box calculations with those from the SM, which are obtained by integration (summation) over momentum. We first discuss the on-shell limit where the nucleon spectral function is represented by a $`\delta `$-function around the bare mass while the $`\mathrm{\Delta }`$ spectral function – as implemented in the transport approach – is given by (in the $`\mathrm{\Delta }`$ rest frame) $$A_\mathrm{\Delta }(M)=\frac{2M^2\mathrm{\Gamma }_{\pi N}(M)}{(M^2M_{\mathrm{\Delta }0}^2)^2+M^2\mathrm{\Gamma }_{\pi N}^2(M)}$$ (45) with $$\mathrm{\Gamma }_{\pi N}(M)=\mathrm{\Gamma }_0(\frac{q}{q_R})^3(\frac{q_R^2+\delta ^2}{q^2+\delta ^2})^3,$$ (46) where $`q_R`$ denotes the pion momentum in the $`\mathrm{\Delta }`$ rest frame of the resonance and $`q`$ is the corresponding pion three-momentum at invariant mass $`M`$, i.e. $$q^2=\frac{(M^2(M_N+M_\pi )^2)(M^2(M_NM_\pi )^2)}{4M^2}.$$ (47) The quantity $`\delta `$ in (46) is fixed by $$\delta ^2=(MM_NM_\pi )^2+\frac{\mathrm{\Gamma }_0^2}{4}$$ (48) with $`\mathrm{\Gamma }_0=120`$ MeV to achieve a good description for the $`\pi N\mathrm{\Delta }`$ reaction in vacuum. Recall that in this case we have $`\mathrm{\Gamma }_{XP}/(2M)=\mathrm{\Gamma }_{decay}=\mathrm{\Gamma }_{\pi N}`$ in the $`\mathrm{\Delta }`$ rest frame, i.e. $`\mathrm{\Gamma }_{coll}`$ = 0. In the upper part of Fig. 3 we compare the asymptotic ($`t\mathrm{}`$) distributions for nucleons ($`N`$) and $`\mathrm{\Delta }`$’s in the on-shell limit from the transport (box) calculation (solid histograms) with the result from the SM at a temperature $`T`$ = 97 MeV employing the $`\mathrm{\Delta }`$ spectral function (45). Since the differences are hardly visible, we conclude that the transport calculation reproduces the result from the SM, where the thermodynamical parameters are determined by energy and baryon number conservation. We mention that we have discarded strangeness in this comparison, since kaons and hyperons are very scarce at this energy and have been switched off in both models. Since the $`\mathrm{\Delta }`$ width (46) is zero below the $`\pi N`$ threshold, the $`\mathrm{\Delta }`$ mass distribution only can extend above $`M_\pi +M_N`$. In case of the off-shell calculation we obtain somewhat different mass distributions for nucleons and $`\mathrm{\Delta }`$’s from the box calculation, which are given in terms of the solid histograms in the lower part of Fig. 3. Here the nucleon spectral function becomes very broad and also the $`\mathrm{\Delta }`$ distribution extends below the $`\pi N`$ threshold in vacuum. The nucleon spectral function is broadened due to the elastic ($`NNNN`$) and inelastic ($`NNN\mathrm{\Delta }`$) scattering processes, which can roughly be described by a collisional width $`\mathrm{\Gamma }_{coll}`$ of 40 MeV in the nucleon spectral function $$A_N(M)=\frac{2M^2\mathrm{\Gamma }_{coll}}{(M^2M_{N0}^2)^2+M^2\mathrm{\Gamma }_{coll}^2},$$ (49) as shown by the left dashed line in the lower part of Fig. 3. The collisional width in case of nucleons is related to $`\mathrm{\Gamma }_{XP}`$ as $`2P_0\mathrm{\Gamma }_{coll}=\mathrm{\Gamma }_{XP}`$ which reduces in the rest system of the particle to $`2M\mathrm{\Gamma }_{coll}=\mathrm{\Gamma }_{XP}`$. The $`\mathrm{\Delta }`$ mass spectrum in this case is more difficult to interprete. This is due to the fact that in the decay $`\mathrm{\Delta }\pi N^{}`$ the $`\mathrm{\Delta }`$ may decay to a pion and an off-shell nucleon $`N^{}`$ according to the nucleon equilibrium spectral function $`A_N(M)`$, which essentially shifts the $`\pi N^{}`$ threshold accordingly. However, taking into account this change in width due to the in-medium $`\pi N^{}`$ decay, only, the low mass part of the $`\mathrm{\Delta }`$ distribution is underestimated considerably. Here the ’collisional’ channels $`\mathrm{\Delta }N\mathrm{\Delta }N`$ and $`\mathrm{\Delta }NNN`$ are much more important. We mention that the latter reaction is described by the extended detailed balance relation of Ref. while the elastic differential cross section is taken the same as for $`NN`$ scattering (as a function of the momentum transfer). Since especially the $`\mathrm{\Delta }N`$ absorption reaction depends on mass $`M`$ and momentum $`p`$ of the baryons explicitly, it is not straightforward to present analytical formulas since final state Pauli blocking for the nucleons leads to a highly nonlinear problem. We thus have extracted the $`\mathrm{\Delta }`$ collisional width $`\mathrm{\Gamma }_{coll}^\mathrm{\Delta }(p,M)`$ from the transport calculation explicitly by calculating the number of $`\mathrm{\Delta }N`$ unblocked collisions per unit time as a function of the $`\mathrm{\Delta }`$ momentum $`p`$ and mass $`M`$. Using ($`p=|\stackrel{}{P}|`$) $$\frac{\mathrm{\Gamma }_{XP}}{2P_0}=\mathrm{\Gamma }_{tot}(p,M)=\mathrm{\Gamma }_{\pi N^{}}(M)+\mathrm{\Gamma }_{coll}^\mathrm{\Delta }(p,M),$$ (50) where $`\mathrm{\Gamma }_{\pi N^{}}`$ denotes the $`\mathrm{\Delta }`$ in-medium $`\pi N^{}`$ width averaged over the nucleon equilibrium distribution function (cf. lower part of Fig. 3) and integrating over momentum with the appropriate Fermi function (for $`T`$ = 97 MeV) we obtain the right dashed line in the lower part of Fig. 3 that describes the box $`\mathrm{\Delta }`$\- distribution rather well. Thus the transport calculation at finite density $`\rho _0`$ and temperature $`T`$ is consistent with the SM when employing the same physical baryon spectral functions. One might have expected this equivalence in equilibrium due to energy and baryon number conservation, however, the actual numerical result then may be regarded as a consistency test of the numerical implementation schemes for the various elastic and inelastic reaction channel in the medium for off-shell particle propagation. The mass integrated number of $`\mathrm{\Delta }`$’s in the off-shell case is larger by a few % as compared to the on-shell case in equilibrium due to the low mass tail of the $`\mathrm{\Delta }`$ distribution, however, the high mass tails of the $`\mathrm{\Delta }`$’s are identical within the statistics achieved as demonstrated in Fig. 4. On the other hand, the number of pions is practically identical in both cases since the low mass tail of the $`\mathrm{\Delta }`$’s is dominantly due to $`\mathrm{\Delta }N`$ reactions and only to a lower extent to the $`\pi N^{}`$ decay as discussed above. ## 4 Summary In this work we have employed the semiclassical off-shell approach from Refs. to study equilibration phenomena in intermediate and high energy nucleus-nucleus collisions in a finite box with periodic boundary conditions. The semiclassical off-shell transport approach describes the virtual propagation of particles in the invariant mass squared $`M^2`$ besides the conventional propagation in the mean-field potential given by the real part of the self energy. The imaginary part of the retarded self energy $`Im\mathrm{\Sigma }_{XP}^{ret}=1/2\mathrm{\Gamma }_{XP}=P_0(\mathrm{\Gamma }_{coll}(XP)+\mathrm{\Gamma }_{decay}(XP))`$ – apart from decay contributions to the width – is determined by the collision integrals themselves and ’evaluated’ within the transport approach dynamically. Our explicit calculations demonstrate that the off-shell propagation of nucleons has practically no sizeable effect on equilibration times especially at lower bombarding energies; more importantly, the off-shell dynamics even lead to a slight increase of the equilibration time for kinetic equilibrium as noticed early by Danielewicz (cf. also Ref. ). Furthermore, we have demonstrated that the off-shell HSD approach reproduces the ’proper’ spectral functions with respect to the statistical model (in case of a grand canonical ensemble) for nucleons and $`\mathrm{\Delta }`$’s that in equilibrium are given analytically once the collisional width $`\mathrm{\Gamma }_{coll}(\stackrel{}{P},M)`$ is known as a function of the 3-momentum $`\stackrel{}{P}`$ and invariant mass $`M`$. The authors like to acknowledge stimulating discussions with E.L. Bratkovskaya, C. Greiner and S. Leupold throughout this study.
warning/0003/gr-qc0003056.html
ar5iv
text
# Fermat-holonomic congruences ## 1 Introduction The relevance of rigid motions in Newtonian mechanics basically stems from the following facts: (i) they model the motions of ideal rigid bodies, and also the behaviour of real rigid bodies at the first approximation level, (ii) they provide a definition of strains which, in elasticity theory, determine stresses, and (iii) they describe the relative motion of two Newtonian reference frames, i.e., those frames where Newton’s laws of mechanics hold, provided that inertial forces are taken into account. A common feature of Newtonian rigid motions is that each one is unambiguously determined by the giving of the trajectory of one point together with the motion’s vorticity along that line (that is, the angular velocity). The relativistic generalization of rigid motions needs to be formulated in terms of a spacetime manifold ($`𝒱_4`$, g). A motion is then defined by a 3-parameter congruence of timelike worldlines, $`𝒞`$, $$x^\alpha (t)=\phi ^\alpha (t,y^1,y^2,y^3)$$ (1) where $`y^1,y^2,y^3`$ are the parameters. In its turn, the congruence $`𝒞`$ is determined by its unit timelike velocity field $$u^\alpha (x),\text{ with}g_{\mu \nu }u^\mu u^\nu =1$$ (2) The stationary space $`_3`$ for the congruence is the quotient space, where cosets are worldlines in $`𝒞`$. The Fermat tensor $$\widehat{g}_{\alpha \beta }:=g_{\alpha \beta }+u_\alpha u_\beta $$ (3) yields the infinitesimal radar distance $$d\widehat{l}_R^2=\widehat{g}_{\mu \nu }(x)dx^\mu dx^\nu $$ (4) between two neighbour worldlines. This quantity is not usually constant along a worldline in $`𝒞`$. Hence, it does not define an infinitesimal distance on $`_3`$. Only in case that the Born-rigidity condition $$\mathrm{\Sigma }_{\alpha \beta }=(u)\widehat{g}_{\alpha \beta }=0$$ (5) holds, $`g_{\alpha \beta }`$ defines a Riemannian metric on $`_3`$. The above condition (5) consists of six independent first order partial differential equations with three independent unknowns, namely, $`u^i`$, $`i=1,2,3`$ (since $`u^4`$ can be obtained from (2)), just like in the Newtonian case. The class of Born-rigid motions would generalize Newtonian rigid motions also because some spatial distance between points in space is conserved. Unfortunately, the Herglotz-Noether theorem states that, even in Minkowski spacetime, the class of Born-rigid motions is narrower than sought. Indeed, motions combining arbitrary acceleration and rotation are excluded from this class. Nevertheless, this shortness should not be surprising. Indeed, six first order partial differential equations for only three unknown functions unavoidably entail integrability conditions, which yield additional equations. The latter will lead to new integrability conditions, and so on. The process of completing the partial differential system (5) ends up with a set of equations which is too restrictive for our expectations (namely, six degrees of freedom or six arbitrary functions of time). In a recent work by one of us, 2-parameter congruences in a (2+1)-dimensional spacetime, $`𝒱_3`$, were considered as a simplification where the condition (5) is still too restrictive (three partial differential equations for two unknowns: $`u^1(x)`$ and $`u^2(x)`$). Then, the vanishing of shear: $$\sigma _{\alpha \beta }\mathrm{\Sigma }_{\alpha \beta }\frac{1}{2}\mathrm{\Sigma }_\mu ^\mu \widehat{g}_{\alpha \beta };\alpha ,\beta =0,1,2.$$ (6) was advanced as a candidate to substitute the condition of Born-rigidity. The condition of vanishing shear (or, equivalently, conformal rigidity) reads: $$\sigma _{\alpha \beta }=0\alpha ,\beta =0,1,2$$ (7) and yields two independent partial differential equations. Indeed, the six equations (7) are constrained by four relations: $$g^{\mu \nu }\sigma _{\mu \nu }=0\mathrm{and}\sigma _{\alpha \mu }u^\mu =0,\alpha =0,1,2.$$ Since the number of unknown functions is also two, the condition (7) can be dealt by standard methods of solution of partial differential systems. Namely, a non-characteristic surface $`𝒮_2𝒱_3`$ and Cauchy data on it must be given so that a unique solution of (7) in a neighbourhood of $`𝒮_2`$ is determined. The surface $`𝒮_2`$ could be, for instance, a 1-parameter subcongruence. In general, the amount of Cauchy data is much larger than our desideratum, namely, one worldline and the vorticity of the congruence in that line. We have however derived a way of getting a congruence out of a part of it. In the particular case that one of the worldlines in the congruence is a geodesic, and the (2+1)-spacetime is flat, reference goes a little further: given the congruence’s vorticity on the geodesic and assuming that strain vanishes on that worldline<sup>1</sup><sup>1</sup>1This condition has been introduced in reference as an enhancement of Einstein’s equivalence principle and named geodesic equivalence principle ., the conformal rigidity condition (7) then determines a unique 2-parameter congruence. The latter would be useful to model a disk whose center is at rest (or in uniform motion), that spins at an arbitrary angular speed, and that remains as rigid as possible. Another remarkable result in is that it exists a flat, rigid, spatial metric, $`\overline{g}_{\alpha \beta }`$, which is conformal to the Fermat tensor, $`\widehat{g}_{\alpha \beta }`$. The fact that the class of conformally rigid congruences in a (2+1)-spacetime is “wide enough” recalls the well know Gauss theorem : > Any Riemannian 2-dimensional space can be conformally mapped into a flat space This suggests us a way to extend the results derived in (2+1)-spacetimes to (3+1)-spacetimes, namely, to inspire the formulation of “meta”-rigidity<sup>2</sup><sup>2</sup>2The word “meta”-rigidity was coined in to generically refer to any relativistic extension of the notion of rigidity conditions in some extension of Gauss theorem to Riemannian 3-manifolds. One instance of the latter is , where Walberer’s theorem is taken as the starting point. In the present paper we shall consider the following ###### Theorem 1 (Riemann) Let (M, g) be a Riemannian 3-manifold. They exist local charts of mutually orthogonal coordinates. Moreover, this can be done in a number of ways. This means that six functions, namely, three coordinates $`y^i`$ and three factors $`f_i`$, $`i=1,2,3`$ can be locally found such that the metric coefficients in this local coordinates are $$g_{ij}(y)=f_i^2(y)\delta _{ij}$$ (8) that is, the Riemannian metric locally admits an orthogonal basis which is holonomic. This result suggests us the following ###### Definition 1 A congruence is said to be Fermat-holonomic iff its Fermat tensor $`\widehat{g}_{\alpha \beta }`$ admits an orthogonal basis which is holonomic. That is, six functions exist: $`y^i(x),`$ $`f_i(x),`$ $`i=1,2,3`$ such that : $$\widehat{g}_{\mu \nu }(x)dx^\mu dx^\nu =f_i^2(x)\delta _{ij}dy^idy^j$$ (9) the summation convention is understood throughout the paper unless the contrary is explicitly indicated (if one of the repeated indices is in brackets, the convention is suspended in that formula). Greek indices run from 1 to 4 and lattin indices from 1 to 3. Section 2 is devoted to develop some geometrical properties of Fermat-holonomic congruences, and in section 3 the existence of these congruence is discussed and posed as a Cauchy problem for a partial differential system. The method is somewhat similar to that used in proving the existence of orthogonal triples of coordinates in a Riemannian 3-manifold (it has been specially inspiring the reading of reference ). In section 4 a given 2-congruence is used as the Cauchy hypersurface for the aforementioned partial differential system, and the problem of getting a Fermat-holonomic congruence out of one of its parts (namely, a 2-parameter subcongruence) is studied. The kinematical meaning of the Cauchy data is also analysed. We must finally insist in the purely local validity of the results here derived. No global aspect of spacetimes has been considered. ## 2 Fermat-holonomic congruences Let $`𝒞`$ be a Fermat-holonomic 3-parameter congruence and let $`u(x)`$ be the unit tangent vector. According to Definition 1, the Fermat tensor, $`\widehat{g}`$, can be written as in equation (9). Consider the differential 1-forms: $$\omega ^i=f_{(i)}dy^i$$ (10) We thus have: $$\widehat{g}=\delta _{ij}\omega ^i\omega ^j$$ (11) Moreover, since $`\widehat{g}`$ is orthogonal to $`u`$, we have that $$i(u)\omega ^l=0$$ (12) As a consequence, the functions $`y^i`$ are constant along any worldline in the congruence: $$u(y^i)=0$$ Let us now introduce the differential 1-form: $$\omega ^4g(u,\mathrm{\_})=u_\alpha (x)dx^\alpha .$$ From (3) and (11) it follows that $$g=\widehat{g}\omega ^4\omega ^4\eta _{\alpha \beta }\omega ^\alpha \omega ^\beta $$ (13) By definition \[equation (10)\] the 1-forms $`\omega ^i`$ must be integrable or, equivalently, they must satisfy: $$d\omega ^i\omega ^i=0$$ (14) As a result we have thus proved the following ###### Theorem 2 Let ($`𝒱_4`$, g) be a spacetime, $`𝒞`$ a Fermat-holonomic 3-parameter congruence and $`u`$ the unit velocity vector. There exist three integrable 1-forms $`\omega ^i`$, i = 1, 2, 3 such that completed with $`\omega ^4g(u,\mathrm{\_})`$, yield a $`g`$-orthonormal basis. The converse theorem can be easily proved too. ###### Theorem 3 If {$`\omega ^\alpha `$}<sub>α=1..4</sub> is a $`g`$-orthonormal thetrad such that $`\omega ^i`$, $`i=1,2,3`$ are spacelike and integrable, then the integral curves of $`u`$, the vector that results from raising the index in $`\omega ^4`$, form a 3-parameter Fermat-holonomic congruence. We shall now present some geometric properties of Fermat-holonomic congruences. ###### Proposition 1 Let $`\omega ^l\mathrm{\Lambda }^1(𝒱_4)`$, $`l=1,2,3`$, be the orthonormal set fulfilling conditions (12), (13) and (14) above. Then $$(u)\omega ^l\omega ^l=0$$ (15) Proof: Using (12) and (14) we can write: $`(u)\omega ^l\omega ^l=[i(u)d\omega ^l+d(i(u)\omega ^l)]\omega ^l=i(u)d\omega ^l\omega ^l`$ $`=i(u)[d\omega ^l\omega ^l]d\omega ^li(u)\omega ^l=0`$ (16) ###### Proposition 2 The strain rate tensor $`\mathrm{\Sigma }(u)\widehat{g}`$ has $`\omega ^i,`$ $`i=1,2,3`$ as principal directions. Furthermore, the same holds for any of its Lie derivatives along the congruence: $`\mathrm{\Sigma }^{(n)}(u)^n\mathrm{\Sigma }=(u)^{n+1}\widehat{g}`$ Proof: As a consequence of proposition 1, they exist three functions $`\varphi _l,`$ $`l=1,2,3`$ such that $`(u)\omega ^l=\varphi _{(l)}\omega ^l`$. Hence $$\mathrm{\Sigma }(u)\widehat{g}=2\varphi _i\delta _{ij}\omega ^i\omega ^j.$$ The second statement, concerning $`\mathrm{\Sigma }^{(n)}`$, is easily shown by induction. $`\mathrm{}`$ A sort of converse result is the following ###### Proposition 3 If $`(u)\mathrm{\Sigma }`$ and $`\mathrm{\Sigma }`$ diagonalize in the same $`g`$-orthonormal basis, then it exists an orthonormal set $`\omega ^l\mathrm{\Lambda }^1(𝒱_4)`$, such that $$(u)\omega ^l\omega ^l=0$$ (17) Proof: According to the hypothesis there exist three 1-forms $`\rho ^i,i=1,2,3`$ such that $`\widehat{g}=\delta _{ij}\rho ^i\rho ^j,`$ $`\mathrm{\Sigma }=2\varphi _i\delta _{ij}\rho ^i\rho ^j`$ (18) $`(u)\mathrm{\Sigma }=2\psi _i\delta _{ij}\rho ^i\rho ^j`$ (19) These $`\rho ^i`$’s are orthogonal to $`u`$, and the same holds for $`(u)\rho ^l`$, hence: $$(u)\rho ^j=A_k^j\rho ^k$$ (20) The latter can be used to calculate the Lie derivatives of (18): $`\mathrm{\Sigma }=(u)\widehat{g}=(A_i^r\delta _{rj}+A_j^r\delta _{ri})\rho ^i\rho ^j`$ $`(u)\mathrm{\Sigma }=2(\dot{\varphi }_i\delta _{ij}+\varphi _i\delta _{ir}A_j^r+\varphi _j\delta _{jr}A_i^r)\rho ^i\rho ^j`$ which compared with (18) and (19) yield: $`(A_i^j+A_j^i)=2\varphi _{(i)}\delta _{ij}`$ (21) $`\dot{\varphi }_{(i)}\delta _{ij}+\varphi _{(i)}A_j^i+\varphi _{(j)}A_i^j)=\psi _{(i)}\delta _{ij}`$ (22) From (21) we have that: $$A_i^i=\varphi _i,A_i^j=A_j^i,ij$$ (23) which substituted in equation (22) yields: $`\text{for }i=j:`$ $`\dot{\varphi _i}+2\varphi _i^2=\psi _i`$ $`\text{for }ij:`$ $`(\varphi _{(i)}\varphi _{(j)})A_i^j=0`$ (24) Now three cases must be considered according to the degeneracy of the eigenvalues of $`\mathrm{\Sigma }`$. * In the case $`\varphi _1\varphi _2\varphi _3`$ the set $`\{\rho ^i\}`$ is unambiguously defined. Eq.(24) implies $`A_i^j=0,ij.`$ Then taking (23) and (20) into account we obtain $`(u)\rho ^i=\varphi _{(i)}\rho ^i`$ and equation (17) follows for $`\omega ^i=\rho ^i`$. * In the case $`\varphi _1=\varphi _2\varphi _3`$ from (23) and (24), we obtain: $$\begin{array}{cc}A_i^i=\varphi _i,\hfill & i=1,2,3;\hfill \\ A_{\mathrm{\hspace{0.33em}2}}^1=A_{\mathrm{\hspace{0.33em}1}}^2\hfill & A_a^3=A_{\mathrm{\hspace{0.33em}3}}^a=0,a=1,2\hfill \end{array}\}$$ (25) which introduced in (20) yield: $$(u)\rho ^3=\varphi _{(3)}\rho ^3;(u)\rho ^a=A_b^a\rho ^ba,b=1,2$$ (26) In the present case, however, the set $$\omega ^3=\rho ^3\text{and }\omega ^a=R_b^a\rho ^ba,b=1,2$$ with $`(R_b^a)O(2)`$ is also a set of eigenvectors for $`\mathrm{\Sigma }`$. Now, an orthogonal matrix $`(R_b^a)`$ can be found such that $`(u)\omega ^a=\varphi \omega ^a`$, with $`\varphi _1=\varphi _2=\varphi `$. Indeed, since $$(u)\omega ^a=[(u)R_b^a(R^1)_c^b+(RAR^1)_c^a]\omega ^c$$ it is enough to require $$(u)R_b^a=R_c^a(A_b^c+\varphi \delta _b^c)$$ which has many solutions $`(R_b^a)O(2)`$ because, by equation (25), $`A_b^c+\varphi \delta _b^c`$ is skewsymmetric. * The completely degenerate case $`\varphi _1=\varphi _2=\varphi _3`$ can be handled in a similar way as case (b). $`\mathrm{}`$ ###### Theorem 4 $`(u)\mathrm{\Sigma }`$ and $`\mathrm{\Sigma }`$ diagonalize in a common $`g`$-orthonormal basis if, and only if, three functions $`A`$, $`B`$ and $`C`$, exist such that $$(u)\mathrm{\Sigma }=A\widehat{g}+B\mathrm{\Sigma }+C\mathrm{\Sigma }^2$$ (27) where $`\mathrm{\Sigma }_{\alpha \beta }^2\mathrm{\Sigma }_{\alpha \mu }\mathrm{\Sigma }_\beta ^\mu `$. Moreover if two among the eigenvalues of $`\mathrm{\Sigma }`$ are equal, then $`C=0`$ can be taken, and in the completely degenerate case, $`B=C=0`$ can be taken. Proof: $`\mathbf{(}\mathbf{}\mathbf{)}`$: Since $`u`$ is orthogonal to both $`\mathrm{\Sigma }`$ and $`(u)\mathrm{\Sigma }`$, we shall have that $`\omega ^4g(u,\mathrm{\_})`$ is in the common orthogonal basis $`\{\omega ^\alpha \}`$. Thus, expressions similar to (18) and (19) hold. Hence to prove (27) amounts to solve the linear system: $$A+2\varphi _iB+4\varphi _i^2C=2\psi _i$$ (28) for the unknowns $`A`$, $`B`$ and $`C`$. The determinant is: $$\mathrm{\Delta }=8(\varphi _2\varphi _1)(\varphi _2\varphi _3)(\varphi _3\varphi _1).$$ In the non-degenerate case $`\mathrm{\Delta }0`$ and (28) has a unique solution. If $`\varphi _1=\varphi _2\varphi _3`$, only the equations for $`l=2`$ and $`3`$ in (28) are independent, there are infinitely many solutions and $`C`$ can be arbitrarily chosen. In particular, $`C=0`$. Finally, in the completely degenerate case, (28) has rank 1, hence it admits infinitely many solutions, and $`B`$ and $`C`$ are arbitrary. $`\mathbf{(}\mathbf{}\mathbf{)}`$: Assume that (27) holds. Since $`\mathrm{\Sigma }`$ diagonalize in a $`g`$-orthonormal basis, we substitute (18) in (27) and it follows immediately that $`(u)\mathrm{\Sigma }`$ diagonalize in the same $`g`$-orthonormal basis. ###### Theorem 5 If $`\mathrm{\Sigma }`$, $`(u)\mathrm{\Sigma }`$ and $`(u)^2\mathrm{\Sigma }`$ diagonalize in a common $`g`$-orthonormal basis, then $`\widehat{g}`$, $`\mathrm{\Sigma }`$, $`(u)\mathrm{\Sigma }`$ and $`(u)^2\mathrm{\Sigma }`$ are linearly dependent. (Hence, the congruence is non-generic .) Proof: By theorem 4, they exist $`A`$, $`B`$ and $`C`$ such that (27) holds. taking the Lie derivative on both sides, using that $`\mathrm{\Sigma }=(u)\widehat{g}`$ and equation (27) itself, and taking into account that the minimal polynomial for $`\mathrm{\Sigma }_\beta ^\alpha `$ has at most degree 3, we arrive at: $$(u)^2\mathrm{\Sigma }=A^{}\widehat{g}+B^{}\mathrm{\Sigma }+C^{}\mathrm{\Sigma }^2$$ (29) where $`A^{}`$, $`B^{}`$ and $`C^{}`$ are some suitable functions. If $`C=0`$, then (27) already proves the theorem. If, on the contrary $`C0`$, we can derive $`\mathrm{\Sigma }^2`$ from (27) and substitute it into (29), so arriving at: $$(u)^2\mathrm{\Sigma }=\left(A^{}\frac{A}{C}\right)\widehat{g}+\left(B^{}\frac{B}{C}\right)\mathrm{\Sigma }+\frac{C^{}}{C}(u)\mathrm{\Sigma }$$ which ends the proof. $`\mathrm{}`$ ## 3 Existence of Fermat-holonomic congruences According to the theorems 2 and 3 in section 2, proving the existence of Fermat-holonomic congruences is equivalent to prove the existence of a $`g`$-orthonormal basis $`\{\omega ^\alpha \}`$ such that: * $`\omega ^i`$ is spacelike, and * $`d\omega ^i\omega ^i=0`$ (14) The dual thetrad will be denoted $`\{e_\alpha \}`$ and the commutation relations: $$[e_\alpha ,e_\beta ]=C_{\alpha \beta }^\mu e_\mu d\omega ^\alpha =\frac{1}{2}C_{\mu \nu }^\alpha \omega ^\mu \omega ^\nu $$ (30) using the latter, (14) can be written as: $$\frac{1}{2}C_{\mu \nu }^i\omega ^\mu \omega ^\nu \omega ^i=0i=1,2,3$$ (31) which in turn is equivalent to: $$\begin{array}{cc}C_{jk}^i=0\hfill & ijk\hfill \\ C_{4j}^i=0\hfill & ij\hfill \end{array}\}$$ (32) Taking into account the relationship between $`C_{\alpha \beta }^\gamma `$ and the Riemannian connexion coefficients in an orthonormal frame , equations (32) are equivalent to: $`\gamma _{jk}^i=0`$ $`ijk`$ (33) $`C_{4j}^i=\gamma _{4j}^i\gamma _{j4}^i=0`$ $`ij`$ (34) For a Riemannian connexion in an orthonormal frame, it holds: $$\gamma _{\alpha k}^i=\gamma _{\alpha i}^k,\gamma _{\alpha k}^4=\gamma _{\alpha 4}^k\alpha =1\mathrm{}4,i,k=1,2,3$$ where the signature $`(+++)`$ has been used. Hence, at most three among the equations (33) are independent. Since $$\gamma _{\alpha \beta }^\gamma =(\omega ^\gamma ,e_\alpha ^{}e_\beta )=\eta ^{\gamma \nu }g(e_\nu ,e_\alpha ^{}e_\beta )$$ (35) equations (33) and (34) yield a first-order partial differential system of nine equations where the unknown is the orthonormal frame $`\{e_\alpha \}`$. ### 3.1 The Cauchy problem Given a hypersurface $`𝒮_0`$, consider a 1-parameter family of hypersurfaces, $`𝒮_\lambda `$, containing it. Let $`n`$ be the unit orthogonal vector field which, by construction, is hypersurface orthogonal, i. e., $`\nu =g(n,\mathrm{\_})`$ is an integrable 1-form. Relatively to $`n`$, each vector $`e_\alpha `$ of the sought frame can be decomposed in an orthogonal part $`e_\alpha ^{}`$ (which is tangent to the hypersurfaces $`𝒮_\lambda `$) and a parallel part, namely: $$e_\alpha =e_\alpha ^{}+n_\alpha n$$ (36) where $`n_\alpha g(e_\alpha ,n)`$. Substituting this decomposition in (33) and (34), taking into account equation (35) and the fact that the connexion is Riemannian, we respectively arrive at: $`\gamma _{jk}^i=n_jW_{ik}+g(e_i,e_j^{}e_k)=0`$ $`ijk`$ (37) $`C_{4k}^i=n_4W_{ik}n_kW_{i4}+H_{ik}=0`$ $`ik`$ (38) where $$W_{\alpha \beta }g(e_\alpha ,n^{}e_\beta )$$ (39) is skewsymmetric, and $$H_{ik}g(e_i,[e_4^{},e_k^{}])+n_i\left(e_4^{}(n_k)e_k^{}(n_4)\right)+n_kg(e_i,e_4^{}n)n_4g(e_i,e_k^{}n)$$ (40) only depends on the unknowns and their derivatives along directions that are tangential to $`𝒮_\lambda `$. If, and only if, $`n_j0`$, the nine equations (37) and (38) can be solved for the six independent components of $`W_{\alpha \beta }`$. A straight manipulation yields: $`W_{ik}={\displaystyle \frac{1}{n_j}}g(e_i,e_j^{}e_k)`$ $`ijk`$ (41) $`W_{i4}={\displaystyle \frac{1}{n_k}}H_{ik}{\displaystyle \frac{n_4}{n_kn_j}}g(e_i,e_j^{}e_k)`$ $`ijk`$ (42) Notice that for each value of $`i=1,2,3`$ there are two ways of choosing $`jk`$ in equation (42). This comes from the fact that the system (37)–(38) is overdetermined. Hence, three subsidiary conditions follow: $$S_i\frac{1}{n_k}H_{ik}\frac{1}{n_j}H_{ij}\frac{n_4}{n_kn_j}g(e_i,e_j^{}e_ke_k^{}e_j)=0ijk$$ (43) these conditions only depend on “tangential derivatives” of the unknowns. ###### Lemma 1 Given an orthonormal thetrad $`\{\overline{e}_\alpha \}`$ on $`𝒮_0`$, such that $`g(n,\overline{e}_i)0`$, it exists a neighbourhood $`𝒰`$ of $`𝒮_0`$ and an orthonormal thetrad $`\{e_\alpha \}`$ which is a solution of $$\begin{array}{c}W_{ik}=\frac{1}{n_j}g(e_i,e_j^{}e_k)\hfill \\ W_{i4}=\frac{1}{n_k}H_{ik}\frac{n_4}{n_kn_j}g(e_i,e_j^{}e_k)\hfill \end{array}\}$$ (44) \[where $`(ikj)`$ is a cyclic permutation of $`(123)`$\] and such that $`e_\alpha =\overline{e}_\alpha `$ on $`𝒮_0`$. Proof: Let $`\{\stackrel{~}{e}_\alpha \}`$ be a given orthonormal frame. In terms of it the unknown frame $`\{e_\alpha \}`$ can be written as: $$e_\alpha =L_\alpha ^\mu \stackrel{~}{e}_\mu $$ where $`(L_\alpha ^\beta )`$ is a Lorentz matrix valued function. Substituting the latter into equation (39) we have: $$W_{\alpha \beta }=L_\alpha ^\mu g(\stackrel{~}{e}_\mu ,n^{}(L_\beta ^\nu \stackrel{~}{e}_\nu )).$$ Then, after a straightforward calculation, we arrive at: $$nL_\alpha ^\beta =L_{}^{\beta \mu }W_{\mu \alpha }\eta ^{\beta \mu }L_\alpha ^\rho g(\stackrel{~}{e}_\mu ,n^{}\stackrel{~}{e}_\rho )$$ (45) where indices are raised by contraction with $`\eta ^{\alpha \beta }`$ and $`nL_\alpha ^beta`$ is the directional derivative along $`n`$. The second term in the right hand side is known and, since $`g(n,\stackrel{~}{e}_i)0`$ on $`𝒮_0`$, equation (44) yield $`W_{\alpha \beta }`$ on $`𝒮_0`$ as a function of $`e_\alpha `$ and their “tangential” derivatives (i. e. L$`{}_{}{}^{\beta }{}_{\alpha }{}^{}`$ and their “tangential” derivatives). Hence, the Cauchy-Kowalevski theorem can be invoked to end the proof. $`\mathrm{}`$ ###### Lemma 2 Let $`\{e_\alpha \}`$ be an orthonormal thetrad, which is a solution of (44) and fulfills the subsidiary conditions (43) on $`𝒮_0`$. Then (43) holds in the neighbourhood $`𝒰`$ of $`𝒮_0`$ where $`\{e_\alpha \}`$ is defined. Proof: Since $`\{e_\alpha \}`$ is a solution of (44), $$C_{jk}^i=0,C_{4k}^i=0C_{4j}^i=n_jS_i$$ (46) where $`(ikj)`$ is a cyclic permutation of $`(123)`$ and $`S_i`$ is defined in (43). The commutation coefficients $`C_{\alpha \beta }^\mu `$ satisfy the Jacobi like identity: $$\eta ^{\alpha \rho \mu \nu }(e_\rho C_{\mu \nu }^\beta C_{\sigma \rho }^\beta C_{\mu \nu }^\sigma )=0$$ (47) Taking $`\alpha =\beta =i`$ in (47) and using (46) we arrive at: $$e_kC_{4j}^i+C_{4j}^i(C_{k(j)}^{(j)}+C_{k4}^4C_{k(i)}^{(i)})=0,\epsilon _{ikj}=1$$ and taking the decomposition (36) into account, and the fact that $`n_k0`$ we have $$n\left(C_{4j}^i\right)+\frac{1}{n_k}[e_k^{}C_{4j}^i+A_kC_{4j}^i]=0,$$ (48) with $`A_k=C_{k(j)}^{(j)}+C_{k4}^4C_{k(i)}^{(i)}`$. The latter can be considered as a partial differential system on $`C_{4j}^i`$. In case that $`n_k0`$, the hypersurface $`𝒮_0`$ is non-characteristic, and the Cauchy-Kowalevski theorem states that (48) has a unique solution, which for $`S_i=0`$ on $`𝒮_0`$, ensures that $`C_{4j}^i=0`$ in a neighbourhood of $`𝒮_0`$ and, equivalently, $`S_i=0`$ in a neighbourhood of $`𝒮_0`$. Summarizing, we have shown the following ###### Theorem 6 Given an orthonormal thetrad $`\{\overline{e}_\alpha \}`$ on $`𝒮_0`$, such that: * $`g(n,\overline{e}_i)0`$ and * $`S_i=0`$ on $`𝒮_0`$ it exists an orthonormal frame $`\{e_\alpha \}`$ in a neighbourhood $`𝒰`$ of $`𝒮_0`$, such that: * it is a solution of (44) and (43), and * $`e_\alpha =\overline{e}_\alpha `$ on $`𝒮_0`$ Thus, the congruence generated by $`u=e_4`$ is Fermat-holonomic. The proof is straightforward from lemmas 1 and 2. $`\mathrm{}`$ ## 4 Getting a Fermat-holonomic 3-congruence out of a given 2-congruence Let $`𝒮_0`$be the hypersurface spanned by a given 2-congruence of worldlines and let $`\overline{u}`$ be the unit velocity vector. (Hereafter, a bar over a symbol indicates that we are only considering the values of that object on $`𝒮_0`$.) ### 4.1 Are the subsidiary conditions consistent? We are interested in completing an orthonormal thetrad $`\{\overline{e}_\alpha \}`$ on $`𝒮_0`$, such that: $`\overline{e}_4=\overline{u}`$, $`\overline{n}_ig(\overline{e}_i,\overline{n})0`$, and that $`S_i=0`$ on $`𝒮_0`$, where $`\overline{n}`$ is the unit vector orthogonal to $`𝒮_0.`$ For each $`\overline{e}_i`$, we consider the decomposition (36): $$\overline{e}_i=\overline{e}_i^{}+\overline{n}_i\overline{n}$$ and take the unit vector $$\widehat{a}_i\frac{\overline{e}_i^{}}{\left|\overline{e}_i^{}\right|}$$ (49) Furthermore, we can consider the combinations: $$\overline{b}_i=ϵ_i^{jk}\overline{n}_j\overline{e}_ki=1,2,3$$ (50) and define the unit vector $`\widehat{b}_i\overline{b}_i/|\overline{b_i}|`$. It is straightforward to see that $`\{\overline{u},\overline{n},\widehat{b}_i,\widehat{a}_i\}`$ is an orthonormal thetrad at any point in $`𝒮_0`$. We also have that $$g(\widehat{a}_i,\widehat{a}_j)=\frac{1}{\left|\overline{e}_i^{}\right|\left|\overline{e}_j^{}\right|}g(\overline{e}_i^{},\overline{e}_j^{})=\frac{1}{\left|\overline{e}_i^{}\right|\left|\overline{e}_j^{}\right|}(\delta _{ij}\overline{n}_i\overline{n}_j)$$ Now, since $`\overline{n}_i0`$, then $`\left|\overline{n}_j\right|<1`$ and it follows that: $$ij\left|g(\widehat{a}_i,\widehat{a}_j)\right|<1\mathrm{and}g(\widehat{a}_i,\widehat{a}_j)0$$ (51) Thus, for any triad $`\{\overline{e}_1,\overline{e}_2,\overline{e}_3\}`$ in $`T_x𝒱_4`$, $`x𝒮_0`$, such that: $`g(\overline{e}_i,\overline{u})=0`$, and $`g(\overline{e}_i,\overline{n})=\overline{n}_i0`$, we can obtain a triad $`\widehat{a}_i`$ $`T_x𝒮_0`$, $`i=1,2,3`$, such that $`g(\widehat{a}_i,\overline{u})=0`$ and that (51) holds. The converse result is the following ###### Theorem 7 Given a triad $`\widehat{a}_i`$ $`T_x𝒮_0`$, $`i=1,2,3`$, such that $`g(\widehat{a}_i,\overline{u})=0`$ and fulfills (51), a triad $`\{\overline{e}_i\}`$ can be obtained such that $`g(\overline{e}_i,\overline{n})0`$ and that $`\{\overline{e}_1,\overline{e}_2,\overline{e}_3,\overline{e}_4=\overline{u}\}`$ is an orthonormal frame. Proof: We must find $`A_i`$ and $`B_i`$ such that $$\overline{e}_i=A_i\widehat{a}_i+B_i\overline{n}$$ since $`\widehat{a}_i`$, and $`\overline{n}`$ are $`u`$-orthogonal, so will be $`\overline{e}_i`$. Now, the condition $`g(\overline{e}_i,\overline{e}_j)=\delta _{ij}`$ implies: $$A_i^2+B_i^2=1$$ (52) $$A_iA_jg(\widehat{a}_i,\widehat{a}_j)+B_iB_j=0ij$$ (53) From which we easily obtain: $$B_i=A_i\sqrt{\frac{g(\widehat{a}_i,\widehat{a}_j)\dot{g}(\widehat{a}_i,\widehat{a}_k)}{g(\widehat{a}_j,\widehat{a}_k)}}ijk$$ (54) and $$A_i=\sqrt{\frac{g(\widehat{a}_j,\widehat{a}_k)}{g(\widehat{a}_j,\widehat{a}_k)g(\widehat{a}_i,\widehat{a}_j)g(\widehat{a}_i,\widehat{a}_k)}}ijk$$ (55) The right hand side of (54) is well defined because, by the hypothesis, the inequalities (51) hold. The denominator in (55) neither vanishes, as a consequence of (51) too. Indeed, denoting by $`\phi _{ij}`$ the angle between $`\widehat{a}_i`$ and $`\widehat{a}_j`$, and taking into account that $`\left|\phi _{ij}\right|+\left|\phi _{jk}\right|+\left|\phi _{ki}\right|=2\pi `$, $`ijk`$ this denominator reads: $$\mathrm{cos}(\phi _{jk})\mathrm{cos}(\phi _{ij})\mathrm{cos}(\phi _{ik})=\mathrm{sin}(\phi _{ij})\mathrm{sin}(\phi _{ik})ijk$$ which does not vanish because $`\left|g(\widehat{a}_i,\widehat{a}_j)\right|=\left|\mathrm{cos}(\phi _{ij})\right|<1,ij`$. $`\mathrm{}`$ We shall attempt to find $`\widehat{a}_1,\widehat{a}_2,\widehat{a}_3`$ such that the triad $`\{\overline{e}_i\}`$ so reconstructed fulfill the subsidiary condition (43). It is easily seen that $`S_i`$ can be written as: $$S_i\frac{1}{n_j}g(e_i,[e_4,e_j])\frac{1}{n_k}g(e_i,[e_4,e_k])=\frac{1}{n_jn_k}g(e_i,[e_4,b_i])$$ (wher $`(ikj)`$ is a cyclic permutation of $`(123)`$) on the hypersurface $`𝒮_0`$, that is: $$\overline{S}_i=\frac{1}{\overline{n}_j\overline{n}_k}g(\overline{e}_i,[\overline{u},\overline{b}_i])$$ Taking now into account that $`\overline{u}`$and $`\overline{b}_i`$ are tangent to $`𝒮_0`$, (hence, $`[\overline{u},\overline{b}_i]`$ is tangent too), we can write: $$\overline{n}_j\overline{n}_k\overline{S}_i=g(\overline{e}_i^{},[\overline{u},\overline{b}_i])=\left|\overline{e}_i^{}\right|\left|\overline{b}_i\right|g(\widehat{a}_i,[\overline{u},\widehat{b}_i])$$ so that the subsidiary conditions $`\overline{S}_i=0`$ are equivalent to $$g(\widehat{a}_i,[\overline{u},\widehat{b}_i])=0$$ (56) Let $`\overline{v}`$ and $`\overline{m}`$ be two unit vector fields such that $`\{\overline{u},\overline{m},\overline{v}\}`$ is an orthonormal basis on the tangent space of $`𝒮_0`$. In terms of this basis, we can write: $$\begin{array}{c}\widehat{b}_i=\mathrm{cos}\theta _i\overline{m}+\mathrm{sin}\theta _i\overline{v}\hfill \\ \widehat{a}_i=\mathrm{sin}\theta _i\overline{m}+\mathrm{cos}\theta _i\overline{v}\hfill \end{array}\}$$ (57) and the conditions (56) lead to: $$\dot{\theta }+\frac{1}{4}\mathrm{sin}2\theta _i(\overline{\mathrm{\Sigma }}_{vv}\overline{\mathrm{\Sigma }}_{mm})+\frac{1}{2}\mathrm{cos}2\theta _i\overline{\mathrm{\Sigma }}_{mv}+\frac{1}{2}\left(g(\overline{v},[\overline{u},\overline{m}])g(\overline{m},[\overline{u},\overline{v}])\right)=0$$ (58) which is an ordinary differential equation for each $`\theta _i`$, $`i=1,2,3`$, and has a solution for every initial data $`\theta _i^0`$ given in a submanifold $`𝒮_0`$, such that it is nowhere tangent to $`\overline{u}`$. Summarizing, we have thus proved that given a 2-congruence, it spans a hypersurface $`𝒮_0`$ on which Cauchy data $`\{\overline{e}_1,\overline{e}_2,\overline{e}_3,\overline{e}_4=\overline{u}\}`$ can be found fulfilling the subsidiary conditions (43). Moreover, this can be done in an infinite number of ways. ### 4.2 The kinematical meaning of the Cauchy data In the particular case considered in this section, where the Cauchy hypersurface $`𝒮_0`$ is spanned by a 2-congruence, we shall analyse the kinematical significance of the Cauchy data $`\{\overline{e}_i\}_{i=1,2,3}`$. According to Proposition 2, the latter is a principal basis for $`\overline{\mathrm{\Sigma }}`$ (i. e., the values of the strain rate of the Fermat-holonomic 3-congruence. We shall see how $`\{\overline{e}_i\}_{i=1,2,3}`$ determine $`\overline{\mathrm{\Sigma }}`$. Indeed, let $`\overline{n}`$ be the unit vector normal to $`𝒮_0`$ and $`\overline{\nu }g(\overline{n},\mathrm{\_})`$. $`\overline{\mathrm{\Sigma }}`$ can be written as: $$\overline{\mathrm{\Sigma }}=\overline{\mathrm{\Sigma }}^0+\overline{\pi }\overline{\nu }+\overline{\nu }\overline{\pi }+\overline{s}\overline{\nu }\overline{\nu }$$ (59) where: $$\overline{\mathrm{\Sigma }}^0(\overline{n},\mathrm{\_})=\overline{\mathrm{\Sigma }}^0(\mathrm{\_},\overline{n})=0\mathrm{and}\overline{\pi },\overline{n}=0.$$ The condition that $`\overline{e}_i`$ is a principal vector then reads: $$\lambda _i\text{such that}\overline{\mathrm{\Sigma }}(\overline{e}_i,\mathrm{\_})=\lambda _i\overline{\omega }^i$$ which, using (59) and considering the parallel and orthogonal parts, respectively yields: $`\overline{\pi },\overline{e}_i+\overline{s}\overline{n}_i=\lambda _i\overline{n}_i`$ (60) $`\overline{\mathrm{\Sigma }}^0(\overline{e}_i^{},\mathrm{\_})+\overline{n}_i\overline{\pi }=\lambda _i\widehat{g}(\overline{e}_i^{},\mathrm{\_})`$ (61) Now, taking into account (36) and expressions like $$\overline{n}^l\overline{e}_l=\overline{n}\text{and}\underset{l=1}{\overset{3}{}}(\overline{n}^l)^2=1,$$ after a short calculation, we arrive at: $`\overline{p}={\displaystyle \underset{jl}{}}{\displaystyle \frac{1}{\overline{n}_j}}\overline{\mathrm{\Sigma }}^0(\overline{e}_j^{},\overline{e}_l^{})\overline{e}_l^{},`$ $`\text{where}\overline{\pi }=\widehat{g}(\overline{p},\mathrm{\_})`$ (62) $`\overline{s}=2\overline{\mathrm{\Sigma }}^0(\overline{e}_j^{},\overline{e}_l^{})\delta ^{lj}+{\displaystyle \underset{j<l}{}}{\displaystyle \frac{1}{\overline{n}_j\overline{n}_l}}\overline{\mathrm{\Sigma }}^0(\overline{e}_j^{},\overline{e}_l^{})`$ (63) $`\lambda _i=\overline{\mathrm{\Sigma }}^0(\overline{e}_h^{},\overline{e}_l^{})\delta ^{lh}+{\displaystyle \frac{1}{\overline{n}_j\overline{n}_k}}\overline{\mathrm{\Sigma }}^0(\overline{e}_j^{},\overline{e}_k^{})`$ (64) Let us now see what does $`\overline{\mathrm{\Sigma }}^0`$ mean. Given any couple of vector fields, $`\overline{v}`$ and $`\overline{w}`$, that are tangent to $`𝒮_0`$, we have that: $`\overline{\mathrm{\Sigma }}^0(\overline{v},\overline{w})=`$ $`\overline{\mathrm{\Sigma }}(\overline{v},\overline{w})=(u)\widehat{g}(\overline{v},\overline{w})`$ $`=`$ $`\overline{u}\left(\widehat{g}(\overline{v},\overline{w})\right)(\widehat{g}([\overline{u},\overline{v}],\overline{w})(\widehat{g}(\overline{v},[\overline{u},\overline{w}])`$ Now, since $`𝒮_0`$ is a submanifold, $`[\overline{u},\overline{v}]`$ and $`[\overline{u},\overline{w}]`$ are also tangent to $`𝒮_0`$. Hence, the Fermat tensor $`\widehat{g}`$ on the right hand side can be replaced by its restriction to $`𝒮_0`$, namely, $`\widehat{g}^0`$. So that, $$\overline{\mathrm{\Sigma }}^0=(\overline{u})\widehat{g}^0$$ Thus, $`\overline{\mathrm{\Sigma }}^0`$ is the strain rate of the given 2-congruence, $`𝒞_2`$ in the Riemannian submanifold $`(𝒮_0,\widehat{g}^0)`$. We have so far shown the following ###### Proposition 4 The Cauchy data $`(𝒮_0,\overline{e}_1,\overline{e}_2,\overline{e}_3,\overline{e}_4=\overline{u})`$ determine $`\overline{\mathrm{\Sigma }}`$, i. e., the values on $`𝒮_0`$ of the strain rate of the Fermat-holonomic congruence obtained as a solution of the partial differential system (37)–(38). The converse is true only if all the eigenvalues of $`\overline{\mathrm{\Sigma }}`$ restricted to spacelike vectors have multiplicity 1. Indeed, if some eigenvalue of $`\overline{\mathrm{\Sigma }}`$ has multiplicity greater than 1, then the eigenvectors are determined up to a rotation. ### 4.3 Uniqueness From Proposition 4 it seems to follow that, given the 2-congruence $`𝒞_2`$ and $`\overline{\mathrm{\Sigma }}`$, there is a unique Fermat-holonomic 3-congruence that includes $`𝒞_2`$ and its strain rate takes the values $`\overline{\mathrm{\Sigma }}`$ on the submanifold $`𝒮_0`$. We shall now see that, although this is the generic case, it is not allways true. Once $`𝒞_2`$ is given, $`\overline{\mathrm{\Sigma }}`$ is no more arbitrary. Indeed, the first block $`\overline{\mathrm{\Sigma }}^0`$ in the decomposition (59) is already determined by $`𝒞_2`$. Hence, it is enough to give the 1-form $`\overline{\pi }`$ and the scalar $`\overline{s}`$ on $`𝒮_0`$, and use (59) to determine $`\overline{\mathrm{\Sigma }}`$. Then, a basis $`\{\overline{e}_1,\overline{e}_2,\overline{e}_3\}`$ of spacelike eigenvectors of $`\overline{\mathrm{\Sigma }}`$ can be chosen. In order that $`(𝒮_0,\overline{e}_1,\overline{e}_2,\overline{e}_3,\overline{e}_4=\overline{u})`$ is a suitable set of Cauchy data for the partial differential system (37)–(38) it is necessary that $`\overline{n}_i0`$, $`i=1,2,3`$. This is equivalent to require that no $`\overline{e}_l`$ is tangent to $`𝒮_0`$. Now, for $`\overline{e}_l`$ to be tangent to $`𝒮_0`$ we should have that: $$\overline{\pi },\overline{e}_l=0\text{and}\overline{\mathrm{\Sigma }}^0(\overline{e}_l,\mathrm{\_})=\lambda _l\widehat{g}^0(\overline{e}_l,\mathrm{\_}),$$ where (60) and (61) have been used. That is, $`\overline{e}_l`$ is a principal vector of $`\overline{\mathrm{\Sigma }}^0`$ and $`\overline{\pi }`$ is orthogonal to $`\overline{e}_l`$. Hence, $`\overline{p}`$ itself should be an eigenvector of $`\overline{\mathrm{\Sigma }}^0`$. Thus, avoiding to choose $`\overline{p}`$ among the eigenvectors of $`\overline{\mathrm{\Sigma }}^0`$ guarantees that $`\overline{n}_i0`$, $`i=1,2,3`$. Of course, this is not possible when $`\overline{\mathrm{\Sigma }}^0`$ is a multiple of $`\widehat{g}^0`$. So that, several possibilities occur: * Either $`\overline{\mathrm{\Sigma }}`$ has three simple eigenvalues and therefore the eigenvectors $`\{\overline{e}_i\}_{i=1,2,3}`$ are uniquely determined, in which case: + either $`\overline{n}_i0`$, $`i=1,2,3`$, and $`(𝒮_0,\overline{e}_1,\overline{e}_2,\overline{e}_3,\overline{e}_4=\overline{u})`$ is a suitable set of Cauchy data, + or some $`\overline{n}_l=0`$, and $`𝒮_0`$ is characteristic. * Or some eigenvalue of $`\overline{\mathrm{\Sigma }}`$ is not simple. Then, the basis of eigenvectors is determined up to a rotation. Therefore, only in the case (A.1) $`𝒞_2`$ and $`\overline{\mathrm{\Sigma }}`$ determine a unique Fermat-holonomic 3-congruence. ## 5 Conclusion and outlook Looking for a less restrictive substitute for Born’s relativistic definition of rigid motion, we have suggested the definition of Fermat-holonomic motion, namely, a 3-parameter congruence of timelike worldlines which admits an adapted system of coordinates $`(t,y^1,y^2,y^3)`$, such that the hypersurfaces $`y^i=`$constant, $`i=1,2,3`$, are mutually orthogonal (i.e., the Fermat tensor is diagonal: $`\widehat{g}_{ij}(t,y)=0`$ whenever $`ij`$). We have expressed this condition as a partial differential system and analysed the Cauchy problem. We have proved —theorem 6— that, given a 2-parameter congruence $`𝒞_2`$ and a triad of spatial vectors $`\{\overline{e}_i\}_{i=1,2,3}`$ on $`𝒮_0`$ (the track of $`𝒞_2`$), there is a unique Fermat-holonomic 3-congruence, $`𝒞_3`$, containing $`𝒞_2`$ and admitting an adapted coordinate system such that the spatial coordinate curves are tangent to $`\overline{e}_i`$ on $`𝒮_0`$. Moreover, the given directions $`\{\overline{e}_i\}`$ are related to the eigenvectors of the strain rate tensor $`\overline{\mathrm{\Sigma }}`$ and, except in the case that the shear of $`𝒞_3`$ vanishes on $`𝒮_0`$, we have shown that $`𝒞_2`$ together with $`\overline{\mathrm{\Sigma }}`$ determine $`𝒞_3`$ in a neigbourhood of $`𝒮_0`$. Hence, a Fermat-holonomic congruence is determined by “a part of it”. The definition of Fermat-holonomic congruences has been devised as an extension to a (3+1)-spacetime of the shear-free congruences studied in reference in a similar context, for the simplified problem of a (2+1)-spacetime. There, we went further and, adding symmetry arguments and the so called geodesic equivalence principle , , a unique shear-free congruence was obtained out of a worldline and the angular velocity on it, just like in the case of Newtonian rigid motions. In a forthcoming paper we shall try to supplement in a similar way (symmetries plus geodesic equivalence principle) the general results that have been derived here, to model an arbitrary rotational motion with a fixed point. ## 6 Acknowledgments One of us is indebted to M. A. Garcia Bonilla for helpful suggestions and comments. This work is partly supported by DIGICyT, contract no.PPB96-0384 and by Institut d’Estudis Catalans (S.C.F.).
warning/0003/hep-th0003200.html
ar5iv
text
# References 1. The quotient conditions. The $`P_{}=N/R`$ sector of the discrete light–cone quantization of uncompactified M–theory is given by the supersymmetric quantum mechanics of $`U(N)`$ matrices. In temporal gauge, the action reads $$S=\frac{1}{2R}𝑑t\mathrm{Tr}\left(\dot{X}^\mu \dot{X}_\mu +\underset{\mu >\nu }{}[X^\mu ,X^\nu ]^2+i\mathrm{\Theta }^T\dot{\mathrm{\Theta }}\mathrm{\Theta }^T\mathrm{\Gamma }_\mu [X^\mu ,\mathrm{\Theta }]\right),$$ (1) where $`\mu ,\nu =1,\mathrm{},9`$. The compactification of M(atrix) theory as a model for M–theory has been studied in . In it has been treated using noncommutative geometry . These investigations apply to the $`d`$–dimensional torus $`T^d`$, and have been further dealt with from various viewpoints in . These structures are also relevant in noncommutative string and gauge theories . Let $`e_{ij}`$, $`i,j=1,2`$, generate a 2-dimensional lattice in $`R^2`$. In compactifying M(atrix) theory on the torus $`T^2`$ determined by this lattice one introduces unitary operators $`𝒰_1`$ and $`𝒰_2`$, defined on the covering space $`R^2`$ of $`T^2`$, such that $`𝒰_i^1X_j𝒰_i`$ $`=`$ $`X_j+2\pi e_{ij},i,j=1,2,`$ (2) $`𝒰_i^1X_a𝒰_i`$ $`=`$ $`X_a,a=3,\mathrm{},9`$ (3) $`𝒰_i^1\mathrm{\Theta }𝒰_i`$ $`=`$ $`\mathrm{\Theta }.`$ (4) By consistency the operators $`𝒰_1`$ and $`𝒰_2`$ commute, up to a constant phase: $$𝒰_1𝒰_2=e^{2\pi i\theta }𝒰_2𝒰_1.$$ (5) In this paper we extend Eqs.(4)(5) to the case of compact Riemann surfaces of genus $`g>1`$. This is a first step towards the compactification of M(atrix) theory on a Riemann surface. The explicit solutions and their supersymmetry properties will be considered elsewhere. A Riemann surface $`\mathrm{\Sigma }`$ of genus $`g>1`$ is constructed as the quotient $`H/\mathrm{\Gamma }`$, where $`H`$ is the upper half–plane, and $`\mathrm{\Gamma }\mathrm{PSL}_2(R)`$, $`\mathrm{\Gamma }\pi _1(\mathrm{\Sigma })`$, is a Fuchsian group acting on $`H`$ as $$\gamma =\left(\begin{array}{c}a\\ c\end{array}\begin{array}{cc}b& \\ d& \end{array}\right)\mathrm{\Gamma },\gamma z=\frac{az+b}{cz+d}.$$ (6) In the absence of elliptic and parabolic generators, the $`2g`$ Fuchsian generators $`\gamma _j`$ satisfy $$\underset{j=1}{\overset{g}{}}\left(\gamma _{2j1}\gamma _{2j}\gamma _{2j1}^1\gamma _{2j}^1\right)=I.$$ (7) Inspired by M(atrix) theory, let us promote the complex coordinate $`z=x+iy`$ to an $`N\times N`$ complex matrix $`Z=X+iY`$, with $`X=X^{}`$ and $`Y=Y^{}`$. This would suggest defining fractional linear transformations of $`Z`$ through conjugation $`𝒰Z𝒰^1=(aZ+bI)(cZ+dI)^1`$. However, taking the trace we see that this construction cannot be implemented for finite $`N`$. Thus we will consider some suitable modification. For the moment note that requiring the $`𝒰_k`$ to represent the $`\gamma _k`$ gives $$\underset{k=1}{\overset{g}{}}\left(𝒰_{2k1}𝒰_{2k}𝒰_{2k1}^1𝒰_{2k}^1\right)=e^{2\pi i\theta }I,$$ (8) which generalizes the relation of the noncommutative torus (5). 2. The noncommutative torus revisited. In order to compactify in higher genus it is necessary to extract some general guidelines from the case of the torus. In $`g=1`$ the fundamental group is Abelian. This implies that the associated differential generators commute, i.e. $`[_1,_2]=0`$, so it makes sense to apply the Baker–Campbell–Hausdorff (BCH) formula when computing the phase $`e^{2\pi i\theta }`$. On the contrary, the fundamental group of negatively curved Riemann surfaces is nonabelian, and the BCH formula is not useful. The derivation of the phase in $`g=1`$ by means of techniques alternative to the BCH formula will be the key point to solving the problem in $`g>1`$. Mimicking the case of $`g=1`$, one expects the building blocks for the solution to the quotient conditions in $`g>1`$ to have the form $`e^{_n_n^{}}`$ or $`e^{i(_n+_n^{})}`$, for some gauged $`\mathrm{sl}_2(R)`$ operators $`_n`$ to be determined. We will show that finding such $`_n`$ is closely connected with the computation of the phase without using the BCH formula. In $`g=1`$ the BCH formula is useful, as the commutator between covariant derivatives can be a constant. On the contrary, in $`g>1`$, the $`_n`$ will be a sort of gauged $`\mathrm{sl}_2(R)`$ generators, and $`[_n,_m]`$ can never be a c–number. The solution to the quotient conditions in $`g=1`$ is expressed in terms of the exponential of covariant derivatives $`_k`$, $`k=1,2`$, so apparently we should use both $`e^{_n_n^{}}`$ and $`e^{i(_n+_n^{})}`$ when passing to $`g>1`$. While the exponential $`e^_k`$ will generate translations, the operator $`e^{_n_n^{}}`$ will produce $`\mathrm{PSL}_2(R)`$ transformations. As the latter are real, we are forced to discard $`e^{i(_n+_n^{})}`$ and to use $`e^{_n_n^{}}`$ only. This fact is strictly related to the nonabelian nature of the group $`\pi _1(\mathrm{\Sigma })`$. Let us consider the operators $$𝒰_k=e^{\lambda _k(_k+iA_k)},\lambda _kR,k=1,2,$$ (9) We also introduce the functions $`F_k(x_1,x_2)`$ defined by $$𝒰_k=F_ke^{\lambda _k_k}F_k^1,k=1,2.$$ (10) The identity $`Af(B)A^1=f(ABA^1)`$ and Eq.(10) give $`(_k+iA_k)F_k=0`$. Also note that $`e^{\lambda _k_k}F_k^1(\{x_k\})=F_k^1(\{x_j+\delta _{jk}\lambda _k\})e^{\lambda _k_k}`$. Therefore, defining $`G_k(x_1,x_2)`$ by $$𝒰_k=G_ke^{\lambda _k_k},k=1,2,$$ (11) we conclude that $$F_k(\{x_j+\delta _{jk}\lambda _k\})=G_k^1(\{x_j\})F_k(\{x_j\}).$$ (12) The unitary operators $`𝒰_k`$ can be used to derive the phase of Eq.(5). First we note that pulling the derivatives to the right we get $$𝒰_1𝒰_2𝒰_1^1𝒰_2^1=F_1e^{\lambda _1_1}F_1^1F_2e^{\lambda _2_2}F_2^1F_1e^{\lambda _1_1}F_1^1F_2e^{\lambda _2_2}F_2^1$$ $$=F_1(x_1,x_2)F_1^1(x_1+\lambda _1,x_2)F_2(x_1+\lambda _1,x_2)F_2^1(x_1+\lambda _1,x_2+\lambda _2)$$ $$\times F_1(x_1+\lambda _1,x_2+\lambda _2)F_1^1(x_1,x_2+\lambda _2)F_2(x_1,x_2+\lambda _2)F_2^1(x_1,x_2).$$ (13) Let us consider the curvature of $`A=A_1dx_1+A_2dx_2`$ $$F=dA=(_1A_2_2A_1)dx_1dx_2=F_{12}dx_1dx_2.$$ (14) The constant–curvature connection is the unique possible choice to get a constant phase, so we set $`F_{12}=2\pi \theta `$. To be explicit we pick the gauge $`A_1=\pi \theta x_2`$, $`A_2=\pi \theta x_1`$, so that $`𝒰_1=e^{i\pi \lambda _1\theta x_2}e^{\lambda _1_1}`$, $`𝒰_2=e^{i\pi \lambda _2\theta x_1}e^{\lambda _2_2}`$, and $$G_1=e^{i\pi \lambda _1\theta x_2}=e^{i\lambda _1A_1},G_2=e^{i\pi \lambda _2\theta x_1}=e^{i\lambda _2A_2},$$ (15) so Eq.(12) reads $$F_1(x_1+\lambda _1,x_2)=e^{i\pi \lambda _1\theta x_2}F_1(x_1,x_2),F_2(x_1,x_2+\lambda _2)=e^{i\pi \lambda _2\theta x_1}F_2(x_1,x_2).$$ (16) The solution is $`F_1=e^{i\pi \theta x_1x_2}f_1(x_2)`$, $`F_2=e^{i\pi \theta x_1x_2}f_2(x_1)`$, with $`f_1`$ ($`f_2`$) an arbitrary function of $`x_2`$ ($`x_1`$). Substituting this into (13) we get Eq.(5), as we would using BCH. From (11)(15) one would understand that the connection in (9) can be simply pulled to the left. However, this is the case only if one chooses a particular gauge, as in general we have $`e^{\lambda _k(_k+iA_k)}e^{i\lambda _kA_k}e^{\lambda _k_k}`$. Indeed, under the gauge transformation $`A_kA_k+_k\chi `$, we have $$e^{\lambda _k(_k+iA_k+i_k\chi )}=e^{i\chi }e^{\lambda _k(_k+iA_k)}e^{i\chi }=e^{i\chi (\{x_j+\delta _{jk}\lambda _j\})i\chi (\{x_j\})}e^{\lambda _k(_k+iA_k)},$$ (17) whereas under a gauge transformation, $`e^{i\lambda _kA_k}e^{\lambda _k_k}`$ is multiplied by $`e^{i\lambda _k\chi (\{x_j\})}`$. It is easily seen that the correct expression is $$e^{\lambda _k(_k+iA_k)}=e^{i_{x_k}^{x_k+\lambda _k}𝑑a_kA_k}e^{\lambda _k_k},$$ (18) where in the integrand one has $`A_1(a_1,x_2)`$ if $`k=1`$ and $`A_2(x_1,a_2)`$ if $`k=2`$. In (18) we used a shorthand notation; the integration limits should be written more precisely as $`_{\{x_j\}}^{\{x_j+\delta _{jk}\lambda _j\}}`$. In particular, the contour is easily recognized as the path joining $`x_k`$ and $`x_k+\lambda _k`$ along the line with $`x_{jk}`$ fixed. Since on the torus we can choose the zero curvature metric, straight lines correspond to geodesics of the metric. Thus, the above contour is the geodesic joining $`\{x_j\}`$ with $`\{x_j+\delta _{jk}\lambda _k\}`$. A direct check of Eq.(18) is that $$e^{\lambda _k(_k+iA_k)}=e^{i_{x_k^0}^{x_k}𝑑a_kA_k}e^{\lambda _k_k}e^{i_{x_k^0}^{x_k}𝑑a_kA_k}=e^{i_{x_k}^{x_k+\lambda _k}𝑑a_kA_k}e^{\lambda _k_k},$$ (19) where we used the property that $`_k_{x_{k0}}^{x_k}𝑑a_kA_k=A_k(x_1,x_2)`$. This is a distinguished feature due to the flatness of the torus that does not hold in $`g>1`$. However, we redefine the contour integral for the torus in a way which easily generalizes to higher genus, namely, The contour integral is along the geodesic, with respect to the constant curvature metric, joining the points with coordinates $`\{x_j\}`$ and $`\{x_j+\delta _{jk}\lambda _k\}`$. Due to the fact that along the integration contour either $`dx_1=0`$ or $`dx_2=0`$, we can replace $`da_kA_k`$ with $`A`$: $$𝒰_k=e^{\lambda _k(_k+iA_k)}=e^{i_{x_k^0}^{x_k}A}e^{\lambda _k_k}e^{i_{x_k^0}^{x_k}A}=e^{i_{x_k}^{x_k+\lambda _k}A}e^{\lambda _k_k},$$ (20) so that $`F_k=e^{i_{x_k^0}^{x_k}A}`$. Even if on the torus the $`F_k`$ are not essential, we introduced them as their higher–genus analog will lead to a new class of functions. By Stokes’ theorem $$𝒰_1𝒰_2𝒰_1^1𝒰_2^1=$$ $$\mathrm{exp}\left[i_{(x_1,x_2)}^{(x_1+\lambda _1,x_2)}A+i_{(x_1+\lambda _1,x_2)}^{(x_1+\lambda _1,x_2+\lambda _2)}A+i_{(x_1+\lambda _1,x_2+\lambda _2)}^{(x_1,x_2+\lambda _2)}A+i_{(x_1,x_2+\lambda _2)}^{(x_1,x_2)}A\right]$$ $$=\mathrm{exp}\left(i_{}A\right)=\mathrm{exp}\left(i_{}F\right)=e^{2\pi i\lambda _1\lambda _2\theta },$$ (21) where $``$ is a fundamental domain for the torus. Note that $`\lambda _1\lambda _2`$ is the area of the torus. Normalizing the area to $`1`$, we get Eq.(5). We now show that the only possible connection leading to a constant value of $`_{}A`$ is the one with constant curvature. In order to denote the dependence on the basepoint of the domain we use the notation $`_{x_1x_2}`$. Independence from $`(x_1,x_2)`$ means that $$_{_{x_1x_2}}F=_{_{x_1^{}x_2^{}}}F,$$ (22) for any $`(x_1^{},x_2^{})R^2`$. Any point in $`R^2`$ can be obtained by a translation $`(x_1,x_2)(x_1^{},x_2^{})=\mu (x_1,x_2)(x_1+b_1,x_2+b_2)`$. Noticing that $`_{x_1^{}x_2^{}}=\mu _{x_1x_2}`$, we see that Eq.(22) is satisfied only if the curvature two–form $`F`$ is invariant under arbitrary translations of $`(x_1,x_2)`$. This fixes $`F`$ to be a constant two–form. The above investigation captures the essence of the construction in $`g=1`$, somehow extracting it from its specific context. This is very useful to reformulate the problem of deriving a projective unitary representation of the fundamental group of a class of manifolds which is much more general than the torus. We can say that in order to get a projective unitary representation of the fundamental group of a given manifold $``$ by means of operators acting on the space $`L^2()`$, we should consider the previous well–defined guidelines. 3. Projective unitary representation of $`\pi _1(\mathrm{\Sigma })`$ on $`L^2(H)`$. We now apply the above general analysis to the case of higher genus Riemann surfaces. We start by first considering a unitary representation of $`\pi _1(\mathrm{\Sigma })`$ realized on $`L^2(H)`$ (the analog of $`e^{\lambda _k_k}`$). For $`n=1,0,1`$ and $`e_n(z)=z^{n+1}`$ let us set $`\mathrm{}_n=e_n(z)_z`$. Define $$L_n=e_n^{1/2}\mathrm{}_ne_n^{1/2}=e_n\left(_z+\frac{1}{2}\frac{e_n^{}}{e_n}\right).$$ (23) They satisfy the $`\mathrm{sl}_2(R)`$ algebra $$[L_m,L_n]=(nm)L_{m+n},[\overline{L}_m,L_n]=0,[L_n,f]=z^{n+1}_zf.$$ (24) For $`k=1,2,\mathrm{},2g`$, consider the operators $$T_k=e^{\lambda _1^{(k)}(L_1+\overline{L}_1)}e^{\lambda _0^{(k)}(L_0+\overline{L}_0)}e^{\lambda _1^{(k)}(L_1+\overline{L}_1)},$$ (25) with the $`\lambda _n^{(k)}`$ picked such that $$T_kzT_k^1=\gamma _kz=\frac{a_kz+b_k}{c_kz+d_k},$$ (26) so that by (7) $$\underset{k=1}{\overset{g}{}}\left(T_{2k1}T_{2k}T_{2k1}^1T_{2k}^1\right)=I.$$ (27) On $`L^2(H)`$ we have the scalar product $`\varphi |\psi =_H𝑑\nu \overline{\varphi }\psi `$, with $`d\nu (z)=idzd\overline{z}/2=dxdy`$. One can check that the $`T_k`$ provide a unitary representation of $`\mathrm{\Gamma }`$. For any function $`F`$ satisfying $`|F|=1`$, we define the operators $$_n^{(F)}=F(z,\overline{z})L_nF^1(z,\overline{z})=e_n\left(_z+\frac{1}{2}\frac{e_n^{}}{e_n}_z\mathrm{ln}F(z,\overline{z})\right),$$ (28) which also satisfy the algebra (24). Its adjoint is given by $$_n^{(F)}=F\overline{e_n^{1/2}}_{\overline{z}}\overline{e_n^{1/2}}F^1=\overline{}_n^{(F^1)}.$$ (29) We now observe that the operators $$\mathrm{\Lambda }_n^{(F)}=_n^{(F)}_n^{(F)}=_n^{(F)}+\overline{}_n^{(F^1)},$$ (30) enjoy the fundamental property that both their chiral components are gauged in the same way by the function $`F`$, that is $$\mathrm{\Lambda }_n^{(F)}=F(L_n+\overline{L}_n)F^1,$$ (31) while also satisfying the $`\mathrm{sl}_2(R)`$ algebra: $$[\mathrm{\Lambda }_m^{(F)},\mathrm{\Lambda }_n^{(F)}]=(nm)\mathrm{\Lambda }_{m+n}^{(F)},[\mathrm{\Lambda }_n^{(F)},f]=(z^{n+1}_z+\overline{z}^{n+1}_{\overline{z}})f.$$ (32) Furthermore, since $`\mathrm{\Lambda }_n^{(F)}=\mathrm{\Lambda }_n^{(F)}`$, the operators $`e^{\mathrm{\Lambda }_n^{(F)}}=Fe^{L_n+\overline{L}_n}F^1`$ are unitary. Let $`b`$ be a real number, and $`A`$ a Hermitean connection to be identified presently. Set $$𝒰_k=e^{ib_z^{\gamma _kz}A}T_k,$$ (33) where the integration contour is taken to be the Poincaré geodesic connecting $`z`$ and $`\gamma _kz`$. As the gauging functions introduced in (28) we will take the $`F_k(z,\overline{z})`$ solutions of the equation $`F_kT_kF_k^1=e^{ib_z^{\gamma _kz}A}T_k`$, that is $$F_k(\gamma _kz,\gamma _k\overline{z})=e^{ib_z^{\gamma _kz}A}F_k(z,\overline{z}).$$ (34) With the choice (34) for $`F_k`$, (31) becomes $$\mathrm{\Lambda }_{n,k}^{(F)}=F_k(L_n+\overline{L}_n)F_k^1=z^{n+1}\left(_z+\frac{n+1}{2z}_z\mathrm{ln}F_k\right)+\overline{z}^{n+1}\left(_{\overline{z}}+\frac{n+1}{2\overline{z}}_{\overline{z}}\mathrm{ln}F_k\right).$$ (35) The $`\mathrm{\Lambda }_{n,k}^{(F)}`$ satisfy the algebra $$[\mathrm{\Lambda }_{m,j}^{(F)},\mathrm{\Lambda }_{n,k}^{(F)}]=(nm)\mathrm{\Lambda }_{m+n,j}^{(F)}+F_k^1|e_n|\mathrm{\Lambda }_{n,k}^{(F)}|e_n|^1F_kF_j^1|e_m|\mathrm{\Lambda }_{m,j}^{(F)}|e_m|^1F_j(\mathrm{ln}F_j\mathrm{ln}F_k),$$ $$[\mathrm{\Lambda }_{n,k}^{(F)},f]=(z^{n+1}_z+\overline{z}^{n+1}_{\overline{z}})f.$$ (36) Upon exponentiating $`\mathrm{\Lambda }_{n,k}^{(F)}`$ one finds $$𝒰_k=e^{\lambda _1^{(k)}\mathrm{\Lambda }_{1,k}^{(F)}}e^{\lambda _0^{(k)}\mathrm{\Lambda }_{0,k}^{(F)}}e^{\lambda _1^{(k)}\mathrm{\Lambda }_{1,k}^{(F)}},$$ (37) that is, the $`𝒰_k`$ are unitary, and $$𝒰_k^1=T_k^1e^{ib_z^{\gamma _kz}A}=e^{ib_{\gamma _k^1z}^zA}T_k^1.$$ (38) It is immediate to see that the $`𝒰_k`$ defined in (33) satisfy (8) for a certain value of $`\theta `$:<sup>1</sup><sup>1</sup>1The differential representation of $`\mathrm{PSL}_2(R)`$ acts in reverse order with respect to the one by matrices. $$\underset{k=1}{\overset{g}{}}\left(𝒰_{2k1}𝒰_{2k}𝒰_{2k1}^{}𝒰_{2k}^{}\right)=e^{ib_z^{\gamma _1z}A}T_1e^{ib_z^{\gamma _2z}A}T_2e^{ib_{\gamma _1^1z}^zA}T_1^1e^{ib_{\gamma _2^1z}^zA}T_2^1\mathrm{}$$ $$=\mathrm{exp}\left[ib\left(_z^{\gamma _1z}+_{\gamma _1z}^{\gamma _2\gamma _1z}+_{\gamma _2\gamma _1z}^{\gamma _1^1\gamma _2\gamma _1z}+_{\gamma _1^1\gamma _2\gamma _1z}^{\gamma _2^1\gamma _1^1\gamma _2\gamma _1z}+\mathrm{}\right)A\right]\underset{k=1}{\overset{g}{}}\left(T_{2k1}T_{2k}T_{2k1}^1T_{2k}^1\right)$$ $$=e^{ib__zA},$$ (39) where $`_z=\{z,\gamma _1z,\gamma _2\gamma _1z,\gamma _1^1\gamma _2\gamma _1z,\mathrm{}\}`$ is a fundamental domain for $`\mathrm{\Gamma }`$. The basepoint $`z`$, plus the action of the Fuchsian generators on it, determine $`_z`$, as the vertices are joined by geodesics. For (39) to provide a projective unitary representation of $`\mathrm{\Gamma }`$, $`__z𝑑A`$ should be $`z`$–independent. Changing $`z`$ to $`z^{}`$ can be expressed as $`zz^{}=\mu z`$ for some $`\mu \mathrm{PSL}_2(R)`$. Then $`_z_{\mu z}=\{\mu z,\gamma _1\mu z,\gamma _2\gamma _1\mu z,\gamma _1^1\gamma _2\gamma _1\mu z,\mathrm{}\}`$. Now consider $`_z\mu _z=\{\mu z,\mu \gamma _1z,\mu \gamma _2\gamma _1z,\mu \gamma _1^1\gamma _2\gamma _1z,\mathrm{}\}`$. The congruence $`\mu _z_{\mu z}`$ follows from two facts: that the vertices are joined by geodesics, and that $`\mathrm{PSL}_2(R)`$ maps geodesics into geodesics. Since $`\mathrm{\Gamma }`$ is defined up to conjugation, $`\mathrm{\Gamma }\mu \mathrm{\Gamma }\mu ^1`$, if $`\mu _z`$ is a fundamental domain, so is $`_{\mu z}`$. Thus, to have $`z`$–independence we need $`\mu \mathrm{PSL}_2(R)`$ $$__z𝑑A=_{_{\mu z}}𝑑A=_{\mu _z}𝑑A=_{}𝑑A.$$ (40) This fixes the (1,1)–form $`dA`$ to be $`\mathrm{PSL}_2(R)`$–invariant. It is well known that the Poincaré form is the unique $`\mathrm{PSL}_2(R)`$–invariant (1,1)–form, up to an overall constant factor. This is a particular case of a more general fact . The Poincaré metric $`ds^2=y^2|dz|^2=2g_{z\overline{z}}|dz|^2=e^\phi |dz|^2`$ has curvature $`R=g^{z\overline{z}}_z_{\overline{z}}\mathrm{ln}g_{z\overline{z}}=1`$, so that $`_{}𝑑\nu e^\phi =2\pi \chi (\mathrm{\Sigma })`$, where $`\chi (\mathrm{\Sigma })=22g`$ is the Euler characteristic. As the Poincaré (1,1)–form is $`dA=e^\phi d\nu `$, this uniquely determines the gauge field to be $$A=A_zdz+A_{\overline{z}}d\overline{z}=\frac{dx}{y},$$ (41) modulo gauge transformations. Using $`_{}A=_{}𝑑A`$ we finally have that (39) becomes $$\underset{k=1}{\overset{g}{}}\left(𝒰_{2k1}𝒰_{2k}𝒰_{2k1}^{}𝒰_{2k}^{}\right)=e^{2\pi ib\chi (\mathrm{\Sigma })}.$$ (42) 4. Nonabelian gauge fields. Up to now we considered the case in which the connection is Abelian. However, it is easy to extend our construction to the nonabelian case in which the gauge group $`U(1)`$ is replaced by $`U(N)`$. The operators $`𝒰_k`$ now become $$𝒰_k=Pe^{ib_z^{\gamma _kz}A}T_k,$$ (43) where the $`T_k`$ are the same as before, times the $`N\times N`$ identity matrix. Eq.(39) is replaced by $$\underset{k=1}{\overset{g}{}}\left(𝒰_{2k1}𝒰_{2k}𝒰_{2k1}^{}𝒰_{2k}^{}\right)=Pe^{ib__zA}.$$ (44) Given an integral along a closed contour $`\sigma _z`$ with basepoint $`z`$, the path–ordered exponentials for a connection $`A`$ and its gauge transform $`A^U=U^1AU+U^1dU`$ are related by $$Pe^{i_{\sigma _z}A}=U(z)Pe^{i_{\sigma _z}A^U}U^1(z)=U(z)Pe^{i_{\sigma _z}𝑑\sigma ^\mu _0^1𝑑ss\sigma ^\nu U^1(s\sigma )F_{\nu \mu }(s\sigma )U(s\sigma )}U^1(z).$$ (45) Applying this to (44), we see that the only possibility to get a coordinate–independent phase is for the curvature (1,1)–form $`F=dA+[A,A]/2`$ to be the identity matrix in the gauge indices times a (1,1)–form $`\eta `$, that is $`F=\eta I`$. It follows that $$Pe^{ib_{}A}=e^{ib_{}F}.$$ (46) This is only a necessary condition for coordinate–independence. However, this is the same as the Abelian case so that $`\eta `$ should be proportional to the Poincaré (1,1)–form. Denoting by $`E`$ the vector bundle on which $`A`$ is defined, we have $`k=\mathrm{deg}(E)=\frac{1}{2\pi }\mathrm{tr}_{}F`$. Set $`\mu (E)=k/N`$ so that $`_{}F=2\pi \mu (E)I`$ and $`\eta =\frac{\mu (E)}{\chi (\mathrm{\Sigma })}e^\phi d\nu `$, i.e. $$F=2\pi \mu (E)\omega I,$$ (47) where $`\omega =\left(e^\phi /_{}𝑑\nu e^\phi \right)d\nu `$. Thus, by (46) we have that Eq.(44) becomes $$\underset{k=1}{\overset{g}{}}\left(𝒰_{2k1}𝒰_{2k}𝒰_{2k1}^{}𝒰_{2k}^{}\right)=e^{2\pi ib\mu (E)}I,$$ (48) which provides a projective unitary representation of $`\pi _1(\mathrm{\Sigma })`$ on $`L^2(H,C^N)`$. 5. Hochschild cohomology and gauge lengths. A basic object is the gauge length function $`d_A(z,w)=_z^wA`$, where the contour integral is along the Poincaré geodesic connecting $`z`$ and $`w`$. In the Abelian case $$d_A(z,w)=_{\mathrm{Re}z}^{\mathrm{Re}w}\frac{dx}{y}=i\mathrm{ln}\left(\frac{z\overline{w}}{w\overline{z}}\right),$$ (49) which is equal to the angle $`\alpha _{zw}`$ spanned by the arc of geodesic connecting $`z`$ and $`w`$. Observe that the gauge length of the geodesic connecting two punctures, i.e. two points on the real line, is $`\pi `$. This is to be compared with the usual divergence of the Poincaré distance. Under a $`\mathrm{PSL}_2(R)`$–transformation $`\mu `$, we have ($`\mu _x_x\mu x`$) $$d_A(\mu z,\mu w)=d_A(z,w)\frac{i}{2}\mathrm{ln}\left(\frac{\mu _z\overline{\mu }_w}{\overline{\mu }_z\mu _w}\right).$$ (50) Therefore, the gauge length of an $`n`$–gon $$d_A^{(n)}(\{z_k\})=\underset{k=1}{\overset{n}{}}d_A(z_k,z_{k+1})=\pi (n2)\underset{k=1}{\overset{n}{}}\alpha _k,$$ (51) where $`z_{n+1}z_1`$, $`n3`$, and $`\alpha _k`$ are the internal angles, is $`\mathrm{PSL}_2(R)`$–invariant. We now show that the length of the triangle is proportional to the Hochschild 2–cocycle of $`\mathrm{\Gamma }`$. The Fuchsian generators $`\gamma _k\mathrm{\Gamma }`$ are projectively represented by means of unitary operators $`𝒰_k`$ acting on $`L^2(H)`$. The product $`\gamma _k\gamma _j`$ is represented by $`𝒰_{jk}`$, which equals $`𝒰_j𝒰_k`$ up to a phase: $$𝒰_j𝒰_k=e^{2\pi i\theta (j,k)}𝒰_{jk}.$$ (52) Associativity implies $$\theta (j,k)+\theta (jk,l)=\theta (j,kl)+\theta (k,l).$$ (53) We can easily determine $`\theta (j,k)`$: $$𝒰_j𝒰_k=\mathrm{exp}\left(ib_z^{\gamma _jz}A+ib_{\gamma _jz}^{\gamma _k\gamma _jz}Aib_z^{\gamma _k\gamma _jz}A\right)𝒰_{jk}=\mathrm{exp}\left(ib_{\tau _{jk}}A\right)𝒰_{jk},$$ (54) where $`\tau _{jk}`$ denotes the geodesic triangle with vertices $`z`$, $`\gamma _jz`$ and $`\gamma _k\gamma _jz`$. This identifies $`\theta (j,k)`$ as the gauge length of the perimeter of the geodesic triangle $`\tau _{jk}`$ times $`b/2\pi `$. By Stokes’ theorem this is the Poincaré area of the triangle. One can check that $`\theta (j,k)`$ in fact satisfies (53). This phase has been considered in different contexts, such as the quantum Hall effect on $`H`$ and Berezin’s quantization of $`H`$ and Von Neumann algebras . The information on the compactification of M(atrix) theory is encoded in the action of $`\mathrm{\Gamma }`$ on $`H`$, plus a projective representation of $`\mathrm{\Gamma }`$. The latter amounts to the choice of a phase. Physically inequivalent choices of $`\theta (j,k)`$ turn out to be in one–to–one correspondence with elements in the 2nd Hochschild cohomology group of $`\mathrm{\Gamma }`$, which is $`U(1)`$. Hence $`\theta =b\chi (\mathrm{\Sigma })`$ is the unique parameter for this compactification ($`\theta =b\mu (E)`$ in the general case). The Poincaré metric is $`\mathrm{PSL}_2(R)`$ invariant whereas $`A`$ is not. So the equality $`_{}A=_{}F`$ should be a consequence of the fact that the variation of $`A`$ under a $`\mathrm{PSL}_2(R)`$ transformation, $`z\mu z=(az+b)/(cz+d)`$, corresponds to a total derivative. In fact we have $$\mathrm{PSL}_2(R):Ai\frac{d\mu z+d\mu \overline{z}}{\mu z\mu \overline{z}}=Ai_z\mathrm{ln}(cz+d)dz+i_{\overline{z}}\mathrm{ln}(c\overline{z}+d)d\overline{z}.$$ (55) Since $`cz+d`$ has no zeroes, we have that $`\mathrm{ln}(cz+d)`$ is a genuine function on $`H`$. It follows that $`i_z\mathrm{ln}(cz+d)dz+i_{\overline{z}}\mathrm{ln}(c\overline{z}+d)d\overline{z}`$, can be written as an external derivative so that Eq.(55) becomes $$\mathrm{PSL}_2(R):AA+d\mathrm{ln}(\mu _z/\overline{\mu }_z)^{\frac{i}{2}},$$ (56) where $`\mu _z_z\mu z`$. So a $`\mathrm{PSL}_2(R)`$–transformation of $`A`$ is equivalent to a gauge transformation. Under $`AA+d\chi `$ we have $`_z^wA_z^wA+\chi (w)\chi (z)`$, which for $`\chi (z)=\mathrm{ln}(\mu _z/\overline{\mu }_z)^{\frac{i}{2}}`$, becomes $$_z^wA_z^wA+\frac{i}{2}\mathrm{ln}\frac{\overline{\mu }_z\mu _w}{\mu _z\overline{\mu }_w}.$$ (57) 6. Preautomorphic forms. Another reason why the gauge–length function is important is that it also appears in the definition (34) of the $`F_k`$. The latter functions, which apparently never appeared in the literature before, are of particular interest. By (34) and (49), $$F_k(\gamma _kz,\gamma _k\overline{z})=\left(\frac{\gamma _kz\overline{z}}{z\gamma _k\overline{z}}\right)^bF_k(z,\overline{z}).$$ (58) Since under a $`\mathrm{PSL}_2(R)`$ transformation the factor $`(w\overline{z})/(z\overline{w})`$ gets transformed by a factor which is typical of automorphic forms, we call the $`F_k`$ preautomorphic forms. Eq.(34) indicates that finding the most general solution to (58) is a problem in geodesic analysis. In the case of the inversion $`\gamma _kz=1/z`$ and $`b`$ an even integer, a solution to (58) is $`F_k=\left(z/\overline{z}\right)^{\frac{b}{2}}`$. By (49) $`F_k=\left(z/\overline{z}\right)^{\frac{b}{2}}`$ is related to the $`A`$–length of the geodesic connecting $`z`$ and $`0`$: $$e^{\frac{i}{2}b_z^0A}=F_k(z,\overline{z})=\left(\frac{z}{\overline{z}}\right)^{\frac{b}{2}}.$$ (59) An interesting formal solution to (58) is $$F_k(z,\overline{z})=\underset{j=0}{\overset{\mathrm{}}{}}\left(\frac{\gamma _k^jz\gamma _k^{j1}\overline{z}}{\gamma _k^{j1}z\gamma _k^j\overline{z}}\right)^b.$$ (60) Consider the uniformizing map $`J_H:H\mathrm{\Sigma }`$, which enjoys the property $`J_H(\gamma z)=J_H(z)`$, $`\gamma \mathrm{\Gamma }`$. Then another solution to (58) is given by $`G(J_H,\overline{J}_H)F_k`$, where $`G`$ is an arbitrary function of the uniformizing map. We should require $`|G|=1`$ for $`|F_k|=1`$. 7. Relation with Donaldson’s approach to stable bundles. We now present some facts about projective, unitary representations of $`\mathrm{\Gamma }`$ and the theory of holomorphic vector bundles . Let $`E\mathrm{\Sigma }`$ be a holomorphic vector bundle over $`\mathrm{\Sigma }`$ of rank $`N`$ and degree $`k`$. The bundle $`E`$ is called stable if the inequality $`\mu (E^{})<\mu (E)`$ holds for every proper holomorphic subbundle $`E^{}E`$. We may take $`N<k0`$. We will further assume that $`\mathrm{\Gamma }`$ contains a unique primitive elliptic element $`\gamma _0`$ of order $`N`$ ($`i.e.`$, $`\gamma _0^N=I`$), with fixed point $`z_0H`$ that projects to $`x_0\mathrm{\Sigma }`$. Given the branching order $`N`$ of $`\gamma _0`$, let $`\rho :\mathrm{\Gamma }U(N)`$ be an irreducible unitary representation. It is said admissible if $`\rho (\gamma _0)=e^{2\pi ik/N}I`$. Putting the elliptic element on the right–hand side, and setting $`\rho _k\rho (\gamma _k)`$, (7) becomes $$\underset{j=1}{\overset{g}{}}\left(\rho _{2j1}\rho _{2j}\rho _{2j1}^1\rho _{2j}^1\right)=e^{2\pi ik/N}I.$$ (61) On the trivial bundle $`H\times C^NH`$ there is an action of $`\mathrm{\Gamma }`$: $`(z,v)(\gamma z,\rho (\gamma )v)`$. This defines the quotient bundle $$H\times C^N/\mathrm{\Gamma }H/\mathrm{\Gamma }\mathrm{\Sigma }.$$ (62) Any admissible representation determines a holomorphic vector bundle $`E_\rho \mathrm{\Sigma }`$ of rank $`N`$ and degree $`k`$. When $`k=0`$, $`E_\rho `$ is simply the quotient bundle (62) of $`H\times C^NH`$. The Narasimhan–Seshadri (NS) theorem now states that a holomorphic vector bundle $`E`$ over $`\mathrm{\Sigma }`$ of rank $`N`$ and degree $`k`$ is stable if and only if it is isomorphic to a bundle $`E_\rho `$, where $`\rho `$ is an admissible representation of $`\mathrm{\Gamma }`$. Moreover, the bundles $`E_{\rho _1}`$ and $`E_{\rho _2}`$ are isomorphic if and only if the representations $`\rho _1`$ and $`\rho _2`$ are equivalent. A differential–geometric approach to stability has been given by Donaldson . Fix a Hermitean metric on $`\mathrm{\Sigma }`$, for example the Poincaré metric, normalized so that the area of $`\mathrm{\Sigma }`$ equals 1. Let us denote by $`\omega `$ its associated (1,1)–form. A holomorphic bundle $`E`$ is stable if and only if there exists on $`E`$ a metric connection $`A_D`$ with central curvature $`F_D=2\pi i\mu (E)\omega I`$; such a connection $`A_D`$ is unique. The unitary projective representations of $`\mathrm{\Gamma }`$ we constructed above have a uniquely defined gauge field whose curvature is proportional to the volume form on $`\mathrm{\Sigma }`$. With respect to the representation considered by NS, we note that NS introduced an elliptic point to produce the phase, while in our case the latter arises from the gauge length. Our construction is directly connected with Donaldson’s approach as $`F=iF_D`$, where $`F`$ is the curvature (47). The main difference is that our operators are unitary differential operators on $`L^2(H,C^N)`$ instead of unitary matrices on $`C^N`$. This allowed us to obtain a non–trivial phase also in the Abelian case. It is however possible to understand the formal relation between our operators and those of NS. To see this we consider the adjoint representation of $`\mathrm{\Gamma }`$ on $`\mathrm{End}C^N`$, $$\mathrm{Ad}\rho (\gamma )Z=\rho (\gamma )Z\rho ^1(\gamma ),$$ (63) where $`Z\mathrm{End}C^N`$ is understood as an $`N\times N`$ matrix. Let us also consider the trivial bundle $`H\times \mathrm{End}C^NH`$. The action of $`\mathrm{\Gamma }`$ $`(z,Z)(\gamma z,\mathrm{Ad}\rho (\gamma )Z)`$ defines the quotient bundle $$H\times \mathrm{End}C^N/\mathrm{\Gamma }H/\mathrm{\Gamma }\mathrm{\Sigma }.$$ (64) Then the idea is to consider a vector bundle $`E^{}`$ in the double scaling limit $`N^{}\mathrm{}`$, $`k^{}\mathrm{}`$, with $`\mu (E^{})=k^{}/N^{}`$ fixed, that is $`\mu (E^{})=b\mu (E)`$. In this limit, fixing a basis in $`L^2(H,C^N)`$, the matrix elements of our operators can be identified with those of $`\rho (\gamma )`$. 8. Noncommutative uniformization. Let us now introduce two copies of the upper half–plane, one with coordinates $`z`$ and $`\overline{z}`$, the other with coordinates $`w`$ and $`\overline{w}`$. While the coordinates $`z`$ and $`\overline{z}`$ are reserved to the operators $`𝒰_k`$ we introduced previously, we reserve $`w`$ and $`\overline{w}`$ to construct a new set of operators. We now introduce noncommutative coordinates expressed in terms of the covariant derivatives $$W=_w+iA_w,\overline{W}=_{\overline{w}}+iA_{\overline{w}},$$ (65) with $`A_w=A_{\overline{w}}=1/(2\mathrm{Im}w)`$, so that $`[W,\overline{W}]=iF_{w\overline{w}}`$, where $`F_{w\overline{w}}=i/[2(\mathrm{Im}w)^2]`$. Let us consider the following realization of the $`\mathrm{sl}_2(R)`$ algebra: $$\widehat{L}_1=w,\widehat{L}_0=\frac{1}{2}(w_w+_ww),\widehat{L}_1=_ww_w.$$ (66) We then define the unitary operators $$\widehat{T}_k=e^{\lambda _1^{(k)}(\widehat{L}_1+\overline{\widehat{L}}_1)}e^{\lambda _0^{(k)}(\widehat{L}_0+\overline{\widehat{L}}_0)}e^{\lambda _1^{(k)}(\widehat{L}_1+\overline{\widehat{L}}_1)},$$ (67) where the $`\lambda _n^{(k)}`$ are as in (25). Set $`𝒱_k=\widehat{T}_k𝒰_k`$. Since the $`\widehat{T}_k`$ satisfy (27), it follows that the $`𝒱_k`$ satisfy (48), times the $`N\times N`$ identity matrix, and $$𝒱_k_w𝒱_k^1=\widehat{T}_k_w\widehat{T}_{k}^{}{}_{}{}^{1}=\frac{a_k_w+b_k}{c_k_w+d_k}.$$ (68) Setting $`W=G_wG^1`$, i.e. $`G=(w\overline{w})^2`$, and using $`Af(B)A^1=f(ABA^1)`$, we see that $$𝒱_kW𝒱_k^1=\widehat{T}_kW\widehat{T}_{k}^{}{}_{}{}^{1}=G(\stackrel{~}{w})\widehat{T}_k_w\widehat{T}_{k}^{}{}_{}{}^{1}G^1(\stackrel{~}{w}),$$ (69) where $$\stackrel{~}{w}=\widehat{T}_kw\widehat{T}_{k}^{}{}_{}{}^{1}=e^{\lambda _0^{(k)}}+2\lambda _1^{(k)}(\widehat{L}_0\lambda _1^{(k)}w)\lambda _1^{(k)2}e^{\lambda _0^{(k)}}(\widehat{L}_1+2\lambda _1^{(k)}\widehat{L}_0\lambda _1^{(k)2}w),$$ (70) and by (68) $$𝒱_kW𝒱_k^1=\widehat{T}_kW\widehat{T}_{k}^{}{}_{}{}^{1}=\frac{a_k\stackrel{~}{W}+b_k}{c_k\stackrel{~}{W}+d_k},$$ (71) where $$\stackrel{~}{W}=_w+G(\stackrel{~}{w})[_wG^1(\stackrel{~}{w})],$$ (72) which differs from $`W`$ by the connection term. Eq.(71) can be seen as representing the noncommutative analog of uniformization. 9. $`C^{}`$–algebra. By a natural generalization of the $`n`$–dimensional noncommutative torus, one defines a noncommutative Riemann surface $`\mathrm{\Sigma }_\theta `$ in $`g>1`$ to be an associative algebra with involution having unitary generators $`𝒰_k`$ obeying the relation (42). Such an algebra is a $`C^{}`$–algebra, as it admits a faithful unitary representation on $`L^2(H,C^N)`$ whose image is norm–closed. Relation (42) is also satisfied by the $`𝒱_k`$. However, while the $`𝒰_k`$ act on the commuting coordinates $`z,\overline{z}`$, the $`𝒱_k`$ act on the operators $`W`$ and $`\overline{W}`$. The latter, factorized by the action of the $`𝒱_k`$ in (71), can be pictorially identified with a sort of noncommutative coordinates on $`\mathrm{\Sigma }_\theta `$. Each $`\gamma I`$ in $`\mathrm{\Gamma }`$ can be uniquely expressed as a positive power of a primitive element $`p\mathrm{\Gamma }`$, primitive meaning that $`p`$ is not a positive power of any other $`p^{}\mathrm{\Gamma }`$ . Let $`𝒱_p`$ be the representative of $`p`$. Any $`𝒱C^{}`$ can be written as $$𝒱=\underset{p\{prim\}}{}\underset{n=0}{\overset{\mathrm{}}{}}c_n^{(p)}𝒱_p^n+c_0I,$$ (73) for certain coefficients $`c_n^{(p)}`$, $`c_0`$. A trace can be defined as $`\mathrm{tr}𝒱=c_0`$. In the case of the torus one can connect the $`C^{}`$–algebras of $`U(1)`$ and $`U(N)`$. To see this one can use ’t Hooft’s clock and shift matrices $`V_1`$, $`V_2`$, which satisfy $`V_1V_2=e^{2\pi i\frac{M}{N}}V_2V_1`$. The $`U(N)`$ $`C^{}`$–algebra is constructed in terms of the $`V_k`$ and of the unitary operators representing the $`U(1)`$ $`C^{}`$–algebra. Morita equivalence is an isomorphism between the two. In higher genus, the analog of the $`V_k`$ is the $`U(N)`$ representation $`\rho (\gamma )`$ considered above. One can obtain a $`U(N)`$ projective unitary differential representation of $`\mathrm{\Gamma }`$ by taking $`𝒱_k\rho (\gamma _k)`$, with $`𝒱_k`$ Abelian. This nonabelian representation should be compared with the one obtained by the nonabelian $`𝒱_k`$ constructed above. In this framework it should be possible to understand a possible higher–genus analog of the Morita equivalence. The isomorphism of the $`C^{}`$–algebras is a direct consequence of an underlying equivalence between the $`U(1)`$ and $`U(N)`$ connection. The $`z`$–independence of the phase requires $`F`$ to be the identity matrix in the gauge indices. This in turn is deeply related to the uniqueness of the connection we found. The latter is related to the uniqueness of the NS connection. We conclude that Morita equivalence in higher genus is intimately related to the NS theorem. Our operators correspond to the $`N\mathrm{}`$ limit of projective unitary representations of $`\mathrm{\Gamma }`$. These operators may be useful in studying the moduli space of M(atrix) string theory . They also play a role in the $`N\mathrm{}`$ limit of QCD as considered in . Finally, let us note that an alternative proposal of noncommutative Riemann surfaces and $`C^{}`$–algebras has been considered in . Acknowledgments. It is a pleasure to thank D. Bellisai, D. Bigatti, M. Bochicchio, G. Bonelli, L. Bonora, U. Bruzzo, R. Casalbuoni, G. Fiore, L. Griguolo, P.M. Ho, S. Kobayashi, I. Kra, G. Landi, K. Lechner, F. Lizzi, P.A. Marchetti, B. Maskit, F. Rădulescu, D. Sorokin, W. Taylor, M. Tonin and R. Zucchini for comments and interesting discussions. G.B. is supported in part by a D.O.E. cooperative agreement DE-FC02-94ER40818 and by an INFN “Bruno Rossi” Fellowship. J.M.I. is supported by an INFN fellowship. J.M.I., M.M. and P.P. are partially supported by the European Commission TMR program ERBFMRX-CT96-0045.
warning/0003/physics0003081.html
ar5iv
text
# THEORY OF COMPLEX SCATTERING LENGTHS ## Abstract We derive a generalized Low equation for the T-matrix appropriate for complex atom-molecule interaction. The properties of this new equation at very low energies are studied and the complex scattering length and effective range are derived. The recent realization of Bose-Einstein condensation (BEC) of ultracold atoms with the accompanying upsurge of theoretical activities have rekindled interest in low energy collisions of atoms and molecules. The subsequent proposals for the creation of ultracold molecular \[1-5\] and hybrid atomic-molecular BEC intensified the above mentioned interest Of particular importance in the above recent developments is the idea of decay of the condensates. In a series of papers, Dalgarno and collaborators \[8-12\] have looked into the idea of using a complex scattering length to represent the low-energy atom-molecule scattering. Implicit in this is the multichannel nature of the collision process: an atom hits a vibrationally excited molecule at extremely low energies. The open inelastic channels are those where the molecule is excited into lower vibrational states. In this sense one has a depletion of the elastic channel. In Ref. , the quenching of $`H_2`$ molecules in collisions with $`H`$ was considered. It was found that, the inelastic cross-sections and the corresponding depletion rate coefficients were very large for high vibrational levels of $`H_2`$. In the above studies, the following form of low-energy S-wave scattering amplitude is used $$f\left(k\right)=\frac{1}{g\left(k^2\right)i\text{ }k}\text{ ,}$$ (1) where k is the wave number related to the center of mass energy of the colliding partners, $`E`$, by $`\frac{\mathrm{}k^2}{2\mu }=E`$, with $`\mu `$ being the reduced mass of the system. The function $`g\left(k^2\right)`$ is even in $`k`$ and is given by the effective range formula. $$g\left(k^2\right)=\frac{1}{a}+\frac{1}{2}r_ok^2\text{ ,}$$ (2) where a is the scattering length and $`r_{}`$ the effective range, both directly related to the interaction. When applied to atom-molecule scattering at very low energies, with the molecules suffering inelastic transitions to lower vibrational states, the scattering length a is taken to be complex, $`a=\alpha i\beta `$, with $`\beta `$ related to the total inelastic cross-section. The question we raise here is the validity of Eq. (1) with a and eventually $`r_{}`$ taken as complex in the case of the elastic scattering with strong coupling to inelastic channels. Of course, an equivalent one-channel description of the elastic scattering can be formulated with the introduction of an appropriate complex optical potential as described by Feshbach . It is therefore legitimate to inquire about the validity of Eq. $`\left(1\right)`$, originally obtained for real potential, if a complex interaction is used . The general structure of the low energy scattering amplitude is also of potentially fundamental importance to very low energy matter interferometry. This method for the obtention of $`f`$ for molecule-molecule scattering has been quite successful at room temperatures . Extension to very low temperatures of this method seems natural and would welcome studies of the type reported here. For the above purpose, it is useful to summarize the elegant derivation of Eq. $`\left(1\right)`$ given by Weinberg . If we denote the interaction by $`V`$ and the free Green’s function by $`G_o^{(+)}\left(E\right)=\left(EH_o+i\epsilon \right)^1`$, then the T-matrix given by the Lippmann-Schwinger equation $`T^{(+)}=V+VG_o^{(+)}T^{(+)}`$, can be written as $`T^{(+)}=V+VG^{(+)}V`$, with the full Green’s function $`G^{(+)}=\left(E+i\epsilon H_oV\right)^1`$. Using the spectral expansion of $`G^{(+)}`$, with the complete set of bound and scattering states $`\{|B,|\mathrm{\Psi }^{(+)}\}`$, we obtain the Low-equation $$\stackrel{}{k}\left|T^{(+)}\left(E\right)\right|\stackrel{}{k}=\stackrel{}{k}\left|V\right|\stackrel{}{k}+\underset{B}{}\frac{\stackrel{}{k}\left|V\right|BB\left|V\right|\stackrel{}{k}}{E+E_B}+d\stackrel{}{k}^{\prime \prime }\frac{T_{\stackrel{}{k}k^{\prime \prime }}^{(+)}\left(E_k\right)\left(T_{\stackrel{}{k}k}^{(+)}\left(E_k\right)\right)^{}}{EE_{k^{\prime \prime }}+i\epsilon }\text{ .}$$ (3) At very low energies relevant for BEC, we seek a solution $`T_{\stackrel{}{k}k}\left(E\right)T\left(E\right)`$ and writing $`\stackrel{}{k}\left|V\right|\stackrel{}{k}=`$ $`\overline{V}`$ we have $$t^{(+)}\left(E\right)=\overline{V}+\underset{B}{}\frac{\left|g_B\right|^2}{E+E_B}+𝑑\stackrel{}{k}^{\prime \prime }\frac{\left|t^{(+)}\left(E_{k^{\prime \prime }}\right)\right|^2}{EE_{k^{\prime \prime }}+i\epsilon }\text{ .}$$ (4) Calculating now $`t^{(+)}\left(E\right)^1t^{()}\left(E\right)^1`$, we find $$t^{(+)}\left(E\right)^1t^{()}\left(E\right)^1=\frac{t^{()}\left(E\right)t^{(+)}\left(E\right)}{t^{()}\left(E\right)\text{ }t^{(+)}\left(E\right)}\text{ .}$$ (5) Since $`t^{()}\left(E\right)=T\left(Ei\epsilon \right)=\left(T^{(+)}\left(E+i\epsilon \right)\right)^{}`$, if $`V`$ is real, we have $`t^{(+)}\left(E\right)^1t^{()}\left(E\right)^1=2ik`$ $`2\pi \frac{2\mu }{\mathrm{}^2}`$ which is just the discontinuity across the positive energy cut in the complex energy plane. Besides the poles in $`t`$, $`\left(\text{zeros in }\left(t\right)^1\right)`$, the only other terms in $`\left(t^{(+)}\right)^1`$ are entire functions of $`WE+i\epsilon `$. Accordingly, with the identification$`f=\frac{1}{2\pi }\frac{2\mu }{\mathrm{}^2}t`$, Eq. (1) follows. We turn next to a complex interaction $`VV^{}`$. The completeness relation now reads $`\underset{B}{}|BB|+d\stackrel{}{k}|\mathrm{\Psi }_\stackrel{}{k}^{(+)}\text{ }\stackrel{~}{\mathrm{\Psi }}_k^{(+)}|`$ where $`|\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}`$ is the dual scattering state which is a solution of the Schrödinger equation with $`V`$ replaced by $`V^{}`$ . Another form of the completeness relation may also be used, $`\underset{B}{}|BB|+d\stackrel{}{k}|\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{()}\text{ }\mathrm{\Psi }_k^{()}|`$,with $`|\mathrm{\Psi }_\stackrel{}{k}^{()}`$ being the physical scattering state with incoming wave boundary condition $`\left(V^{}\text{,}i\epsilon \right)`$ and $`|\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{()}`$ its corresponding dual state $`\left(V\text{,}i\epsilon \right)`$. Thus, the full Green’s function now has the spectral form $$G^{(+)}\left(E\right)=\underset{B}{}\frac{|BB|}{E+E_B}+d\stackrel{}{k}\frac{|\mathrm{\Psi }_\stackrel{}{k}^{(+)}\text{ }\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}|}{EE_k+i\epsilon }\text{ .}$$ (6) Accordingly, Eq. (3) now reads $$\stackrel{}{k}\left|T\right|\stackrel{}{k}=\stackrel{}{k}\left|V\right|\stackrel{}{k}+\underset{B}{}\frac{\stackrel{}{k}\left|V\right|BB\left|V\right|\stackrel{}{k}}{E+E_B+i\epsilon }+d\stackrel{}{k}\frac{\stackrel{}{k}\left|V\right|\mathrm{\Psi }_\stackrel{}{k}^{(+)}\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}\left|V\right|\stackrel{}{k}}{EE_k+i\epsilon }\text{ .}$$ (7) It is clear that the Low equation, Eq. (3), is not valid anymore. However, as we show below Eq. (1) is still valid, with the appropriate generalization of the real function $`g\left(k^2\right)`$ to a complex one . To see this we analyze the matrix element $`\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}\left|V\right|\stackrel{}{k}`$. From the $`LS`$ equation for $`\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}|`$, $$\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}|=\stackrel{}{k}|+\stackrel{}{k}|V\frac{1}{E_kH_{}Vi\epsilon }\stackrel{}{k}|[1+VG^{()}\left(E_k\right)]\text{ .}$$ (8) Thus $`\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}\left|V\right|\stackrel{}{k}=\stackrel{}{k}\left|\stackrel{~}{T}(E_ki\epsilon )\right|\stackrel{}{k}`$, where the unphysical T-matrix $`\stackrel{~}{T}`$ is given by $$\stackrel{~}{T}=V+VG^{()}V\text{ .}$$ (9) Accordingly the T-matrix equation, Eq (7), may be written as $$\stackrel{}{k}\left|T\left(E\right)\right|\stackrel{}{k}=\stackrel{}{k}\left|V\right|\stackrel{}{k}+\underset{B}{}\frac{\stackrel{}{k}\left|V\right|BB\left|V\right|\stackrel{}{k}}{E+E_B+i\epsilon }$$ (10) $$+d\stackrel{}{k}\frac{\stackrel{}{k}\left|T(E)\right|\stackrel{}{k}\stackrel{}{k}\left|\stackrel{~}{T}(E)\right|\stackrel{}{k}}{EE+i\epsilon }\text{ .}$$ (11) A similar equation holds for $`\stackrel{}{k}\left|\stackrel{~}{T}\left(E\right)\right|\stackrel{}{k}`$with $`i\epsilon `$ replaced by $`i\epsilon `$. It is interesting at this point to show the relation between the physical T-matrix element $`\stackrel{}{k}\left|T\left(E\right)\right|\stackrel{}{k}`$ and $`\stackrel{}{k}\left|\stackrel{~}{T}\left(E\right)\right|\stackrel{}{k}`$. This can be done easily following operator manipulations of , and using the relation $`\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}|=\mathrm{\Psi }_\stackrel{}{k}^{(+)}|+\mathrm{\Psi }_\stackrel{}{k}^{(+)}|\left(VV^{}\right)G^{()}\left(E_k\right)`$, Eq. (8), $$\stackrel{}{k}\left|\stackrel{~}{T}\left(E\right)\right|\stackrel{}{k}=\stackrel{}{k}\left|T\left(E\right)\right|\stackrel{}{k}^{}+d\stackrel{}{k}\mathrm{\Psi }_\stackrel{}{k}^{(+)}\left|(VV^{^{}})\right|\mathrm{\Psi }_\stackrel{}{k}^{(+)}S_{\stackrel{}{k}\stackrel{}{k}}^1\text{ ,}$$ (12) where $`S^1`$ is the inverse S-matrix in the elastic channel, $`S_{\stackrel{}{k}\stackrel{}{k}}^1=\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{(+)}|\stackrel{~}{\mathrm{\Psi }}_\stackrel{}{k}^{()}`$, and the diagonal part of the matrix element $`\mathrm{\Psi }_\stackrel{}{k}^{(+)}\left|\left(VV^{}\right)\right|\mathrm{\Psi }_\stackrel{}{k}^{(+)}`$is directly related to the total inelastic scattering cross-section, $`\sigma _{in}`$, viz $$\mathrm{\Psi }_\stackrel{}{k}^{(+)}\left|\left(VV^{}\right)\right|\mathrm{\Psi }_\stackrel{}{k}^{(+)}=2i\frac{E}{k}\sigma _{in}\left(E\right)\text{ .}$$ (13) Eq. (11) explicitly exhibits the connection between $`\stackrel{~}{T}`$ and $`T`$ through the absorptive part of the effective interaction. Now we seek the low energy solution $`\stackrel{}{k}\left|T\right|\stackrel{}{k}t_+\left(E\right)`$ and $`\stackrel{}{k}\left|\stackrel{~}{T}\right|\stackrel{}{k}t_{}\left(E\right)`$ and following the same steps as Weinberg’s , we find immediately, from Eq. (10), with $`f_\pm =\frac{1}{2\pi }\frac{2\mu }{\mathrm{}^2}t_\pm `$, $$f_+^1=g_c\left(k^2\right)ik\text{ ;}$$ (14) $$f_{}^1=g_c\left(k^2\right)+ik\text{ ,}$$ (15) where $`g_c\left(k^2\right)`$ is the complex generalization of $`g\left(k^2\right)`$ of Eq. (1). We turn now to the connection between $`g_c\left(k^2\right)`$ and the low-energy observables. This is most conveniently accomplished by employing the generalized optical theorem $$\frac{4\pi }{k}Imf_+=\sigma _{el}+\sigma _{in}\text{ ,}$$ (16) where $`\sigma _{el}`$ is the total elastic scattering cross section $`4\pi \left|f_+\right|^2`$ and $`\sigma _{in}`$ the total inelastic cross-section. Using (12), we find $$\frac{Im\text{ }g_c\left(k^2\right)}{\left(Re\text{ }g_c\left(k^2\right)\right)^2+\left(Im\text{ }g_c\left(k^2\right)k\right)^2}=\frac{k}{4\pi }\sigma _{in}\text{ .}$$ (17) At $`k=0`$, $`g_c\left(0\right)=\frac{1}{a}`$, where a is the complex scattering length written as $`\alpha i\beta `$. Thus the imaginary part of a, $`\beta `$, is found to be $$\beta =\frac{\left(k\text{ }\sigma _{in}\right)_{k=0}}{4\pi }\text{ ,}$$ (18) an expression also derived in Ref. . Eq. (16) clearly implies that $`\sigma _{in}`$ should go as $`k^1`$ as $`k`$ is lowered, in accordance with Wigner’s law. We go a bit beyond Refs. \[8-12\] and derive a relation between $`\beta `$ and the imaginary part of the effective potential. Since $`\sigma _{in}`$ is given by (for S-wave scattering), Eq. (12) $$\sigma _{in}=\frac{4\pi }{kE}\underset{0}{\overset{\mathrm{}}{}}\left|u\left(r\right)\right|^2\left|ImV\text{ }\left(r\right)\right|\text{ }𝑑r\text{ ,}$$ (19) where $`u\left(r\right)`$ is the S-wave elastic radial wave function, we find $$\beta =\left[\frac{1}{E}\underset{0}{\overset{\mathrm{}}{}}\left|u\left(r\right)\right|^2\left|ImV\text{ }\left(r\right)\right|\text{ }𝑑r\right]_{E0}\text{ .}$$ (20) An equation for the complex effective range, $`r_{}`$, can also be easily derived. Equation (19) is the principle result of this work. It summarizes the following: 1. The coupled-channels calculation aimed to describe the molecular quenching can be recast as an effective one-channel calculation with a complex interaction whose imaginary part account for flux loss. 2. The low-energy behaviour of the scattering amplitude with the complex interaction alluded to above can be conveniently parametrized in terms of complex scattering length and effective range. The message this work conveys is the potential usefulness of constructing the effective complex (optical) interaction for the scattering of ro-vibrational molecules from atoms at low energies. The calculation of a and $`r_{}`$ from knowledge of this potential can be done in a direct and unambiguous way. Acknowledgement Part of this work was done while the author was visiting ITAMP-Harvard. He wishes to thank Prof. Kate Kirby and Dr. H. Sadeghpour for hospitality. He also thanks Drs. N. Balakrishnan and V. Kharchenko for useful discussion. Partial support form the ITAMP-NSF grant and from FAPESP and CNPq is acknowledged.
warning/0003/gr-qc0003093.html
ar5iv
text
# Acceleration-Induced Nonlocal Electrodynamics in Minkowski Spacetime ## 1 Introduction As pointed out by Einstein , in special relativity theory it is assumed that the rate of a fundamental (“ideal”) clock depends on its instantaneous speed and is not affected by its instantaneous acceleration. This is usually called the “clock hypothesis”; see for more recent discussions of this assumption. The decay of elementary particles obeys this hypothesis very well as shown by Eisele for the weak decay of the muon. If we study electrodynamics, for instance, in an accelerated reference frame (see ), then we have to presuppose corresponding hypotheses for the measurement of the electric and magnetic fields, the electric charge, etc. In this way, we arrive at the hypothesis of locality that has been extensively investigated . Replacing the curved worldline of the accelerated observer by its instantaneous tangent vector is reasonable if the intrinsic spacetime scales of the phenomena under consideration are negligibly small compared to the characteristic acceleration scales that determine the curvature of the worldline; otherwise, the past worldline of the observer must be taken into account. This would then result in a nonlocal electrodynamics for accelerated systems. Nonlocal constitutive relations have been studied in the *phenomenological* electrodynamics of continuous media for a long time . In basic field theories, form-factor nonlocality has been the subject of extensive investigations. The main problem with such field theoretical approaches has been that they defy quantization. A review of nonlocal quantum field theories and their insurmountable difficulties has been given by Marnelius . The present work is concerned with a benign form of nonlocality that is induced by the acceleration of the observer. The hypothesis of locality refers directly to acceleration; therefore, one can develop an alternative approach in which the acceleration enters as the decisive quantity. This type of nonlocality, if it refers to time, would involve persistent memory effects. Materials with memory have been extensively studied. However, we are interested in the “material” vacuum – and in this context our paper is devoted to a comparison of two models involving acceleration-induced nonlocality. ## 2 Mashhoon’s model The observational basis of the special theory of relativity generally involves measuring devices that are accelerated; for instance, static laboratory devices on the Earth participate in its proper rotation. The standard extension of Lorentz invariance to accelerated observers in Minkowski spacetime is based on the hypothesis of locality, namely, the assumption that an accelerated observer is locally equivalent to a momentarily comoving inertial observer. The worldline of an accelerated observer in Minkowski spacetime is curved and this curvature depends on the observer’s translational and rotational acceleration scales. The hypothesis of locality is thus reasonable if the curvature of the worldline could be ignored, i.e. if the phenomena under consideration have intrinsic scales that are negligible as compared to the acceleration scales of the observer. The accelerated observer passes through a continuous infinity of hypothetical comoving inertial observers along its worldline; therefore, to go beyond the hypothesis of locality, it appears natural to relate the measurements of an accelerated observer to the class of instantaneous comoving inertial observers. Consider, for instance, an electromagnetic radiation field $`F_{ij}`$ in an inertial frame and an accelerated observer carrying an orthonormal tetrad frame $`e^i{}_{\alpha }{}^{}(\tau )`$ along its worldline. Here $`\tau `$ is its proper time, the Latin indices $`i`$, $`j`$, $`k`$, …, which run from 0 to 3, refer to spacetime coordinates (holonomic indices), while the Greek indices $`\alpha `$, $`\beta `$, $`\gamma `$, …, which run from $`\widehat{0}`$ to $`\widehat{3}`$, refer to (anholonomic) frame indices, and we choose the signature $`(+,,,)`$. The hypothesis of locality implies that the field as measured by the observer is the projection of $`F_{ij}`$ upon the frame of the instantaneously comoving inertial observer, i.e. $$F_{\alpha \beta }(\tau )=F_{ij}e^i{}_{\alpha }{}^{}e_{}^{j}{}_{\beta }{}^{}.$$ (1) On the other hand, measuring the properties of the radiation field would necessitate finite intervals of time and space that would then involve the curvature of the worldline. The most general linear relationship between the measurements of the accelerated observer and the class of comoving inertial observers consistent with causality is $$_{\alpha \beta }(\tau )=F_{\alpha \beta }(\tau )+\underset{\tau _0}{\overset{\tau }{}}K_{\alpha \beta }{}_{}{}^{\gamma \delta }(\tau ,\tau ^{})F_{\gamma \delta }(\tau ^{})d\tau ^{},$$ (2) where $`_{\alpha \beta }`$ is the *field actually measured*, $`\tau _0`$ is the instant at which the acceleration begins and the kernel $`K`$ is expected to depend on the acceleration of the observer. A nonlocal theory of accelerated observers has been developed based on the assumptions that (i) $`K`$ is a convolution-type kernel, i.e. it depends only on $`\tau \tau ^{}`$, and (ii) the radiation field never stands completely still with respect to an accelerated observer. The latter is a generalization of a consequence of Lorentz invariance for inertial observers to all observers. In the space of continuous functions, the Volterra integral equation (2) provides a unique relationship between $`_{\alpha \beta }`$ and $`F_{\alpha \beta }`$. It is possible to express (2) as $$F_{\alpha \beta }(\tau )=_{\alpha \beta }(\tau )+\underset{\tau _0}{\overset{\tau }{}}R_{\alpha \beta }{}_{}{}^{\gamma \delta }(\tau ,\tau ^{})_{\gamma \delta }(\tau ^{})d\tau ^{},$$ (3) where $`R`$ is the resolvent kernel and if $`K`$ is a convolution-type kernel as we have assumed in (i), then so is $`R`$, i.e. $`R=R(\tau \tau ^{})`$. Assumption (ii) then implies that $$R(\tau )=\frac{d\mathrm{\Lambda }(\tau +\tau _0)}{d\tau }\mathrm{\Lambda }^1(\tau _0),$$ (4) where $`R`$ and $`\mathrm{\Lambda }`$ are $`6\times 6`$ matrices and $`\mathrm{\Lambda }`$ is defined by (1) expressed as $`\widehat{F}=\mathrm{\Lambda }F`$ in the six-vector notation. Here $`\widehat{F}`$ denotes the field as referred to the anholonomic frame. This nonlocal theory, which is consistent with all observational data available at present, has been described in detail elsewhere . Fig.1. The path of an observer in space moving with constant angular velocity around the $`z`$-axis for $`\tau >\tau _0`$. It proves interesting to provide a concrete example of the nonlocal relationship (2). Imagine an observer that moves uniformly in the inertial frame along the $`y`$-axis with speed $`c\beta `$ for $`\tau <\tau _0`$ and for $`\tau \tau _0`$ rotates with uniform angular speed $`\mathrm{\Omega }`$ about the $`z`$-axis on a circle of radius $`r`$, $`\beta =r\mathrm{\Omega }/c`$, in the $`(x,y)`$-plane, see Fig.1. In this case, $`e^i_{\widehat{0}}`$ $`=`$ $`\gamma (1,\beta \mathrm{sin}\phi ,\beta \mathrm{cos}\phi ,\mathrm{\hspace{0.25em}0}),`$ $`e^i_{\widehat{1}}`$ $`=`$ $`(0,\mathrm{cos}\phi ,\mathrm{sin}\phi ,\mathrm{\hspace{0.25em}0}),`$ (5) $`e^i_{\widehat{2}}`$ $`=`$ $`\gamma (\beta ,\mathrm{sin}\phi ,\mathrm{cos}\phi ,\mathrm{\hspace{0.25em}0}),`$ $`e^i_{\widehat{3}}`$ $`=`$ $`(0,0,0,\mathrm{\hspace{0.25em}1}),`$ in $`(ct,x,y,z)`$ coordinates with $`\phi =\mathrm{\Omega }(tt_0)=\gamma \mathrm{\Omega }(\tau \tau _0)`$. Here $`\phi `$ is the azimuthal angle in the $`(x,y)`$-plane and $`\gamma `$ is the Lorentz factor. Using six-vector notation, $$(F_{\alpha \beta })\left[\begin{array}{c}\widehat{𝑬}\\ \widehat{𝑩}\end{array}\right],(_{\alpha \beta })\left[\begin{array}{c}𝓔\\ 𝓑\end{array}\right],$$ (6) one can show that with respect to the tetrad frame (2) $`𝓔`$ $`=`$ $`\widehat{𝑬}+{\displaystyle \underset{\tau _0}{\overset{\tau }{}}}\left[𝝎\times \widehat{𝑬}(\tau ^{}){\displaystyle \frac{𝒂}{c}}\times \widehat{𝑩}(\tau ^{})\right]𝑑\tau ^{},`$ (7) $`𝓑`$ $`=`$ $`\widehat{𝑩}+{\displaystyle \underset{\tau _0}{\overset{\tau }{}}}\left[{\displaystyle \frac{𝒂}{c}}\times \widehat{𝑬}(\tau ^{})+𝝎\times \widehat{𝑩}(\tau ^{})\right]𝑑\tau ^{},`$ (8) where $`𝒂`$ is the constant centripetal acceleration of the observer and $`𝝎`$ is its constant angular velocity. These quantities can be expressed with respect to the triad $`e^i_A`$ as $`𝒂=(c\beta \gamma ^2\mathrm{\Omega },\mathrm{\hspace{0.17em}0},\mathrm{\hspace{0.17em}0})`$ and $`𝝎=(0,\mathrm{\hspace{0.17em}0},\gamma ^2\mathrm{\Omega })`$. For an arbitrary accelerated observer, we expect that the relations analogous to (7) and (8) would be much more complicated. Imagine now a general congruence of accelerated observers such that relations similar to (2) and (3) hold for each member of the congruence. The requirement that the electromagnetic field $`F_{ij}`$ (or $`F_{\alpha \beta }`$) satisfy Maxwell’s equations would then imply, via (3), that the field $`_{\alpha \beta }`$ would satisfy certain complicated integro-differential equations, which could then be regarded as the nonlocal Maxwell equations for $`_{\alpha \beta }`$. Instead of this system, we give here a different, but analogous, acceleration-induced nonlocal electrodynamics and study some of its main properties. ## 3 Charge & flux electrodynamics with a new nonlocal ansatz The electrodynamics of charged particles and flux lines, see and the references cited therein, involves the electromagnetic field strength $`F_{\alpha \beta }`$—that is defined via the Lorentz force law and is directly related to the conservation law of magnetic flux—as well as the electromagnetic excitation $`^{\alpha \beta }`$ that is directly related to the electric charge conservation. The corresponding Maxwell equations are metric-free and in Ricci calculus in arbitary frames read (cf. ) $`_{[\alpha }F_{\beta \gamma ]}C_{[\alpha \beta }{}_{}{}^{\delta }F_{\gamma ]\delta }^{}`$ $`=`$ $`0,`$ (9) $`_\beta ^{\alpha \beta }{\displaystyle \frac{1}{2}}C_{\beta \gamma }{}_{}{}^{\alpha }_{}^{\gamma \beta }{\displaystyle \frac{1}{2}}C_{\beta \gamma }{}_{}{}^{\beta }_{}^{\alpha \gamma }`$ $`=`$ $`𝒥^\alpha .`$ (10) Here $`𝒥^\alpha `$ is the electric current and the $`C`$’s are the components of the object of anholonomicity: $$C_{\alpha \beta }{}_{}{}^{\gamma }:=2e^i{}_{\alpha }{}^{}e_{}^{j}{}_{\beta }{}^{}_{[i}^{}e_{j]}{}_{}{}^{\gamma }=C_{\beta \alpha }{}_{}{}^{\gamma }.$$ (11) Ordinarily for vacuum, we would have the constitutive equation $$^{\alpha \beta }=\sqrt{g}g^{\alpha \mu }g^{\beta \nu }F_{\mu \nu }.$$ (12) However, this reformulation of electrodynamics allows for much more general constitutive relations between $`^{\alpha \beta }`$ and $`F_{\alpha \beta }`$. In particular, it is possible to develop a nonlocal *ansatz* based on a generalization of (12) along the lines suggested by Obukhov and Hehl $$^{\alpha \beta }(\tau ,\xi )=\sqrt{g}g^{\alpha \mu }g^{\beta \nu }𝒦_{\mu \nu }{}_{}{}^{\rho \sigma }(\tau ,\tau ^{},\xi )F_{\rho \sigma }(\tau ^{},\xi )d\tau ^{},$$ (13) where the kernel $`𝒦`$ corresponds to the response of the medium and $`\xi ^A`$, $`A=1,2,3`$, are the Lagrange coordinates of the medium. As an alternative to Mashhoon’s model but along the same line of thought, see equation (2), one can develop an acceleration-induced nonlocal constitutive relation in vacuum via equation (13) by using the ansatz, $`^{\alpha \beta }(\tau )`$ $`=`$ $`\sqrt{g}g^{\alpha \mu }g^{\beta \nu }[F_{\mu \nu }(\tau ).`$ (14) $`c{\displaystyle \underset{\tau _0}{\overset{\tau }{}}}[\mathrm{\Gamma }_{0\mu }{}_{}{}^{\rho }(\tau \tau ^{})F_{\rho \nu }(\tau ^{})+.\mathrm{\Gamma }_{0\nu }{}_{}{}^{\rho }(\tau \tau ^{})F_{\mu \rho }(\tau ^{})]d\tau ^{}],`$ where the integral is over the worldline of an accelerated observer in Minkowski spacetime as before. Here the response of the “medium” is simply given by the Levi-Civita connection of the accelerated observer in vacuum and the local constitutive relation (12) is recovered for *inertial* observers. We recall that in an orthonormal frame the connection is equivalent to the anholonomicity, see : $$\mathrm{\Gamma }_{\alpha \beta \gamma }:=g_{\gamma \delta }\mathrm{\Gamma }_{\alpha \beta }{}_{}{}^{\delta }=\frac{1}{2}(C_{\alpha \beta \gamma }+C_{\beta \gamma \alpha }C_{\gamma \alpha \beta })=\mathrm{\Gamma }_{\alpha \gamma \beta }.$$ (15) If we invert (15), we find that $`C_{\alpha \beta \gamma }=2\mathrm{\Gamma }_{[\alpha \beta ]\gamma }`$. In the following, we explore the consequences of the new ansatz (14) for a general accelerated observer in Minkowski spacetime. ## 4 The new ansatz and the accelerating and rotating observer It has been shown in , and the references cited therein, that the orthonormal frame $`e_\alpha `$ of an arbitrary observer with local 3-acceleration $`𝒂`$ and local 3-angular velocity $`𝝎`$ reads $`e_{\widehat{0}}`$ $`=`$ $`{\displaystyle \frac{1}{1+\frac{𝒂}{c^2}\overline{𝒙}}}\left[_{\overline{0}}\left({\displaystyle \frac{𝝎}{c}}\times \overline{𝒙}\right)^{\overline{B}}_{\overline{B}}\right],`$ $`e_A`$ $`=`$ $`_{\overline{A}},`$ (16) where the barred coordinates are the standard normal coordinates adapted to the worldline of the accelerated observer. The coframe $`\vartheta ^\alpha `$ can be computed by inversion. We find $`\vartheta ^{\widehat{0}}`$ $`=`$ $`\left(1+{\displaystyle \frac{𝒂}{c^2}}\overline{𝒙}\right)dx^{\overline{0}}=Ndx^{\overline{0}},`$ $`\vartheta ^A`$ $`=`$ $`dx^{\overline{A}}+\left({\displaystyle \frac{𝝎}{c}}\times \overline{𝒙}\right)^{\overline{A}}dx^{\overline{0}}=dx^{\overline{A}}+N^{\overline{A}}dx^{\overline{0}}.`$ (17) In the $`(1+3)`$-decomposition of spacetime, $`N`$ and $`N^{\overline{A}}`$ are known as lapse function and shift vector, respectively. The frame and the coframe are orthonormal. The metric reads as follows: $`ds^2=\eta _{\alpha \beta }\vartheta ^\alpha \vartheta ^\beta `$ $`=`$ $`\left[\left(1+{\displaystyle \frac{𝒂}{c^2}}\overline{𝒙}\right)^2\left({\displaystyle \frac{𝝎}{c}}\times \overline{𝒙}\right)^2\right]\left(dx^{\overline{0}}\right)^2`$ (18) $`2\left({\displaystyle \frac{𝝎}{c}}\times \overline{𝒙}\right)_{\overline{A}}dx^{\overline{0}}dx^{\overline{A}}\delta _{\overline{A}\overline{B}}dx^{\overline{A}}dx^{\overline{B}},`$ where $`(𝝎\times \overline{𝒙})_{\overline{A}}=ϵ_{\overline{A}\overline{B}\overline{C}}\omega ^{\overline{B}}x^{\overline{C}}`$, $`𝒂=a^{\overline{A}}e_{\overline{A}}`$, and $`a^{\overline{A}}=e_i{}_{}{}^{\overline{A}}a_{}^{i}`$. Starting with the coframe, we can read off the connection coefficients (for vanishing torsion) by using Cartan’s first structure equation $`d\vartheta ^\alpha =\mathrm{\Gamma }_\beta {}_{}{}^{\alpha }\vartheta ^\beta `$ with $`\mathrm{\Gamma }_\beta {}_{}{}^{\alpha }=\mathrm{\Gamma }_{\overline{i}\beta }{}_{}{}^{\alpha }dx^{\overline{i}}`$. By construction, the connection projected in spacelike directions vanishes, since we have spatial Cartesian laboratory coordinates. Thus we are left with the following nonvanishing connection coefficients: $`\mathrm{\Gamma }_{\overline{0}\widehat{0}A}`$ $`=`$ $`\mathrm{\Gamma }_{\overline{0}A\widehat{0}}={\displaystyle \frac{a_A}{c^2}},`$ $`\mathrm{\Gamma }_{\overline{0}AB}`$ $`=`$ $`\mathrm{\Gamma }_{\overline{0}BA}=ϵ_{ABC}{\displaystyle \frac{\omega ^C}{c}}.`$ (19) The first index in $`\mathrm{\Gamma }`$ is holonomic, whereas the second and third indices are anholonomic. If we transform the first index, by means of the frame coefficients $`e^{\overline{i}}_\alpha `$, into an anholonomic one, then we find the totally anholonomic connection coefficients as follows: $`\mathrm{\Gamma }_{\widehat{0}\widehat{0}A}`$ $`=`$ $`\mathrm{\Gamma }_{\widehat{0}A\widehat{0}}={\displaystyle \frac{a_A/c^2}{1+𝒂\overline{𝒙}/c^2}},`$ $`\mathrm{\Gamma }_{\widehat{0}AB}`$ $`=`$ $`\mathrm{\Gamma }_{\widehat{0}BA}={\displaystyle \frac{ϵ_{ABC}\omega ^C/c}{1+𝒂\overline{𝒙}/c^2}}.`$ (20) In general, of course, the translational acceleration $`𝒂`$ and the angular velocity $`𝝎`$ are functions of time. Let us return to (14). If we study the electric sector of the theory, we find, because of (4), $$^{\widehat{0}B}(\tau )=\eta ^{\widehat{0}\widehat{0}}\eta ^{BD}\left[F_{\widehat{0}D}(\tau )c_{\tau _0}^\tau \left(\mathrm{\Gamma }_{0\widehat{0}}{}_{}{}^{C}F_{CD}^{}+\mathrm{\Gamma }_{0D}{}_{}{}^{C}F_{\widehat{0}C}^{}\right)𝑑\tau ^{}\right]$$ (21) or $$𝑫=𝑬+_{\tau _0}^\tau \left[𝝎(\tau \tau ^{})\times 𝑬(\tau ^{})\frac{𝒂(\tau \tau ^{})}{c}\times 𝑩(\tau ^{})\right]𝑑\tau ^{}.$$ (22) Similarly, for the magnetic sector, the corresponding relations read $$\begin{array}{c}^{AB}=\eta ^{AD}\eta ^{BE}[F_{DE}\hfill \\ \hfill c_{\tau _0}^\tau (\mathrm{\Gamma }_{0D}{}_{}{}^{\widehat{0}}F_{\widehat{0}E}^{}+\mathrm{\Gamma }_{0D}{}_{}{}^{C}F_{CE}^{}+\mathrm{\Gamma }_{0E}{}_{}{}^{\widehat{0}}F_{D\widehat{0}}^{}+\mathrm{\Gamma }_{0E}{}_{}{}^{C}F_{DC}^{})d\tau ^{}]\end{array}$$ (23) or $$𝑯=𝑩+_{\tau _0}^\tau \left[𝝎(\tau \tau ^{})\times 𝑩(\tau ^{})+\frac{𝒂(\tau \tau ^{})}{c}\times 𝑬(\tau ^{})\right]𝑑\tau ^{},$$ (24) respectively. Clearly, for constant $`𝒂`$ and $`𝝎`$ our nonlocal relations (22) and (24) are the same as (7) and (8) provided we identify $``$ with $``$, i.e. we postulate that the field actually measured by the accelerated observer is the excitation $``$. This agreement does not extend to the case of *non*uniform acceleration, however, as will be demonstrated in the next section. ## 5 Nonuniform acceleration To show that the new ansatz (14) is different from Mashhoon’s ansatz (2) for the case of nonuniform acceleration even when we identify $``$ with $``$, we proceed via contradiction. That is, let us assume that $`_{\alpha \beta }=_{\alpha \beta }`$ and hence from (22) and (24) $$K(\tau )=\left[\begin{array}{cc}K_𝝎& K_𝒂\\ K_𝒂& K_𝝎\end{array}\right],$$ (25) where $`K_𝝎=𝝎(\tau )𝑰`$ and $`K_𝒂=𝒂(\tau )𝑰/c`$. Here $`I_A`$, $`(I_A)_{BC}=ϵ_{ABC}`$, is a $`3\times 3`$ matrix that is proportional to the operator of infinitesimal rotations about the $`e_A`$-axis. We must now prove that in general $`R(\tau )`$ given by (4) cannot be the resolvent kernel corresponding to $`K(\tau )`$ given by (25). To this end, consider an observer that is accelerated at $`\tau _0=0`$ and note that for kernels of Faltung type in equations (2) and (3) we can write $$\overline{}=(I+\overline{K})\overline{\widehat{F}}\mathrm{and}\overline{\widehat{F}}=(I+\overline{R})\overline{},$$ (26) respectively, where $`\overline{f}(s)`$ is the Laplace transform of $`f(\tau )`$ defined by $$\overline{f}(s):=_0^{\mathrm{}}f(\tau )e^{s\tau }𝑑\tau $$ (27) and $`I`$ is the unit $`6\times 6`$ matrix. Hence, the relation between $`K`$ and $`R`$ may be expressed as $$(I+\overline{K})(I+\overline{R})=I.$$ (28) Fig.2. The acceleration of an observer that is uniformly accelerated only during a finite interval from $`\tau =0`$ to $`\tau =\alpha `$. Imagine now an observer that is at rest on the $`z`$-axis for $`\mathrm{}<\tau <0`$ and undergoes linear acceleration along the $`z`$-axis at $`\tau =0`$ such that $`a(\tau )=g>0`$ for $`0\tau <\alpha `$ and $`a(\tau )=0`$ for $`\tau \alpha `$ (see Fig. 2). That is, the acceleration is turned off at $`\tau =\alpha `$ and thereafter the observer moves with uniform speed $`c\mathrm{tanh}(g\alpha /c)`$ along the $`z`$-axis to infinity. Thus in (25), $`K_𝝎=0`$ and $`K_𝒂=a(\tau )I_3/c`$. On the other hand, one can show that (4) can be expressed in this case as $$R(\tau )=a(\tau )\left[\begin{array}{cc}U& V\\ V& U\end{array}\right],$$ (29) where $`U=J_3\mathrm{sinh}\mathrm{\Theta }`$, $`V=I_3\mathrm{cosh}\mathrm{\Theta }`$, and $`(J_3)_{AB}=\delta _{AB}\delta _{A3}\delta _{B3}`$. Here we have set $`c=1`$ and $$\mathrm{\Theta }(\tau )=\underset{0}{\overset{\tau }{}}a(\tau )𝑑\tau =\{\begin{array}{cc}g\tau ,\hfill & 0\tau <\alpha ,\hfill \\ g\alpha ,\hfill & \tau \alpha .\hfill \end{array}$$ (30) It is now possible to work out (28) explicitly and conclude that for $$X(s):=\overline{a(\tau )\mathrm{sinh}\mathrm{\Theta }},Y(s):=\overline{a(\tau )\mathrm{cosh}\mathrm{\Theta }},Z(s):=\overline{a(\tau )},$$ (31) we must have $$X=YZ,Y=Z(1+X).$$ (32) These relations imply that $$Y(s)=\frac{Z(s)}{1Z^2(s)}.$$ (33) On the other hand, we have $$Z(s)=\underset{0}{\overset{\mathrm{}}{}}a(\tau )e^{s\tau }𝑑\tau =\frac{g}{s}\left(1e^{\alpha s}\right)$$ (34) and $`Y(s)`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}a(\tau )\left(e^\mathrm{\Theta }+e^\mathrm{\Theta }\right)e^{s\tau }𝑑\tau `$ (35) $`=`$ $`{\displaystyle \frac{g}{2}}\left[{\displaystyle \frac{1e^{(sg)\alpha }}{sg}}+{\displaystyle \frac{1e^{(s+g)\alpha }}{s+g}}\right].`$ We consider only the region $`s>g`$ in which $`X(s)`$ and $`Y(s)`$ remain finite for $`\alpha \mathrm{}`$. Comparing (35) with $$\frac{Z}{1Z^2}=\frac{gs(1e^{\alpha s})}{s^2g^2(1e^{\alpha s})^2},$$ (36) we find that, contrary to (33), they do not agree except in the $`\alpha \mathrm{}`$ limit (see Fig.3). Therefore, we conclude that the two models are different if one considers *arbitrary* accelerations. Fig.3. Plot of the functions $`Y(s)`$ and $`W(s):=Z(s)/[1Z^2(s)]`$ for $`\alpha g=2`$. ## 6 Discussion If one rewrites Mashhoon’s nonlocal electrodynamics in the framework of charge & flux electrodynamics in vacuum by substituting the generalization of equation (3) for a congruence of accelerated observers in equations (9)–(12), one finds a rather complicated implicit nonlocal constitutive law. The Maxwell equations expressed in terms of the excitations $`(𝑫,𝑯)`$ and field strengths $`(𝑬,𝑩)`$ remain the same, a fact which is significant since otherwise the conservation laws of electric charge and magnetic flux would be violated. In this paper, we have developed an alternative nonlocal constitutive ansatz within the framework of charge & flux electrodynamics such that the nonlocality is induced by the acceleration of the observer in a similar way as in Mashhoon’s model. An explicit example of nonuniform acceleration has been used to show that the two nonlocal prescriptions discussed here are in general different. Acknowledgments: One of the authors (FWH) would like to thank Bahram Mashhoon and the Department of Physics & Astronomy of the University of Missouri-Columbia for the invitation for a two-month stay and for the hospitality extended to him. This stay has been supported by the VW-Foundation of Hannover, Germany. We are grateful to Yuri Obukhov (Moscow) for highly interesting and critical discussions on nonlocal electrodynamics. Thanks are also due to Guillermo Rubilar (Cologne) for helpful correspondence.
warning/0003/cond-mat0003480.html
ar5iv
text
# Interface Scaling in the Contact Process ## I Introduction Scaling and criticality in nonequilibrium systems continue to be of great interest in statistical physics. Among the various classes of systems that have been subject to intensive study are models of growing interfaces , and absorbing-state phase transitions, typified by directed percolation (DP) . Absorbing-state phase transitions have been linked to self-organized criticality (SOC) , as have driven interface models . The latter connection is established by defining a height variable $`h_i(t)`$ at each site of a sandpile model; at each time interval, the height increases by one unit at each active (toppling) site. It turns out that the interface of the Bak-Tang-Wiesenfeld model is described by an Edwards-Wilkinson equation with columnar noise . In view of the connections between absorbing-state phase transitions, SOC, and surface growth, it is of interest to study the dynamics of the interface representation of a simple model in the DP class. Precise results on the scaling properties of the DP interface representation should prove useful when trying to assign interface models (or height representations of other models, such as sandpiles) to universality classes. In this paper we examine the dynamics of the contact process in dimensions 1 - 3, studying the interface width and the height-height correlation function, as well as the height probability distribution. In Sec. II we define the model and then present a brief discussion of the associated continuum equation, and of a scaling theory. Numerical results are presented in Sec. III. Sec. IV contains a brief summary. ## II Model The contact process (CP) is a simple particle system (lattice Markov process) exhibiting a phase transition to an absorbing (frozen) state at a critical value of the creation rate . This model belongs to the universality class of directed percolation and Reggeon field theory , and is pertinent to models of epidemics , catalysis , and damage spreading , among many others. In the CP each site of the hypercubic lattice $`𝒵^d`$ is either vacant or occupied by a particle. Particles are created at vacant sites at rate $`\lambda n/2d`$, where $`n`$ is the number of occupied nearest-neighbors, and are annihilated at unit rate, independent of the surrounding configuration. The order parameter is the particle density $`\rho `$; the vacuum state, $`\rho =0`$, is absorbing. As $`\lambda `$ is increased beyond $`\lambda _c`$, there is a continuous phase transition from the vacuum to an active state; for $`\lambda >\lambda _c`$, $`\rho (\lambda \lambda _c)^\beta `$ in the stationary state. In one dimension, $`\lambda _c3.297848`$. There are a number of ways (equivalent as regards scaling behavior), of implementing the CP in a simulation algorithm; this work follows the widely used practice of maintaining a list of all occupied sites. In this study the initial condition is always that of all sites occupied. Subsequent events involve selecting (at random) an occupied site x from the $`N_p`$ sites on the list, selecting a process: creation with probability $`p=\lambda /(1+\lambda )`$, annihilation with probability $`1p`$, and, in the case of creation, selecting one of the $`2d`$ nearest neighbors, y, of x. (The creation attempt succeeds if y is vacant). The time increment $`\mathrm{\Delta }t`$ associated with an event is $`1/N_p`$, where $`N_p`$ is the number of occupied sites immediately prior to the event. A trial ends when all the particles have vanished, or at the first event with time $`t_m`$, a preset maximum time. The important scaling laws pertinent to the critical contact process on a lattice of $`L^d`$ sites, starting with all sites occupied are: (1) the mean survival time $`\tau L^{\nu _{||}/\nu _{}}`$, and (2) the average particle density decays as a power law, $`\rho (t)t^\theta `$ for $`1<t<\tau `$. (In practice, the power law is found already for $`t2`$.) In one dimension, $`\nu _{||}/\nu _{}1.5808`$ and $`\theta 0.1595`$ . Occupied sites represent activity, which spreads from site to site. (The absence of any activity corresponds to the absorbing state.) In a growing surface or driven interface, activity corresponds to the motion of the interface. We therefore define the height $`h_i(t)`$ at site $`i`$ as the amount of time (up to time $`t`$) that site $`i`$ has been occupied. In our numerical studies we use a real-valued height (recall that time is not restricted to integer values in our implementation). By keeping a record of the last time $`t_i`$ at which the state of site $`i`$ changed, we are able to evaluate $`h_i(t)`$ at any moment in the simulation. (While the results reported here are for real $`h`$, we find the same scaling properties for integer $`h`$.) The surface $`h_i(t)`$ may be thought of as a driven interface. Since the critical contact process must eventually enter the absorbing state, the interface, in this analogy, will eventually be pinned. The large-scale properties of the CP and related models such as DP can be described via a field theory (so-called Reggeon field theory) framed in terms of a coarse-grained density $`\rho (𝐱,t)0`$ . Retaining only relevant terms, the stochastic pde for $`\rho (𝐱,t)`$ reads $$\frac{\rho }{t}=^2\rho a\rho b\rho ^2+\eta (x,t).$$ (1) Here $`\eta (𝐱,t)`$ is a Gaussian noise with zero mean and autocorrelation $$\overline{\eta (𝐱,t)\eta (𝐱^{},t^{})}=\mathrm{\Gamma }\rho (𝐱,t)\delta ^d(𝐱𝐱^{})\delta (tt^{}).$$ (2) In our continuum description, the height is given by $$h(𝐱,t)=_0^t𝑑t^{}\rho (𝐱,t^{}).$$ (3) Integrating Eq. (1) from time zero to time $`t`$, we obtain $$\frac{h}{t}=^2hahb_0^t\left(\frac{h}{t^{}}\right)^2𝑑t^{}+\zeta (𝐱,t),$$ (4) with the noise autocorrelation $$\overline{\zeta (𝐱,t)\zeta (𝐱^{},t^{})}=\mathrm{\Gamma }\delta ^d(𝐱𝐱^{})h(𝐱,t_<),$$ (5) where $`t_<\mathrm{min}(t,t^{})`$. Thus the equation governing the height includes a nonlinear memory term and a noise with nonvanishing correlations between different times. This equation does not seem to shed much light on the scaling properties of the interface. Rather, the relation between $`h`$ and the density in the contact process, Eq. (3), yields some properties that are not immediately obvious from Eq. (4), for example, that $`h/t0`$, and that the nonlinear term is relevant for $`d<d_c=4`$. In any event, Eq. (4) does serve to point up some differences between the CP interface and conventional surface-growth models such as the Edwards-Wilkinson or Kardar-Parisi-Zhang equations . First, the linear “drive” term ($`ah(𝐱,t)`$ with $`a<0`$ in the active regime) is proportional to the local height and so cannot be transformed away. Second, in the active state, the moduli of the last three terms on the r.h.s. of Eq. (4) grow without limit, suggesting that $`^2h`$ does as well, so that the width of the active phase never saturates. We turn now to a simple scaling analysis of the interface. At its basis lies a scaling hypothesis for the probability density $`p(h;t)`$ of the height $`h`$ (at any lattice site) at time $`t`$: the time-dependence of this density enters only through the mean height $`\overline{h}(t)`$. The conjectured scaling property is $$p(h;t)=\frac{1}{\overline{h}(t)}𝒫(h/\overline{h}(t)),$$ (6) where the scaling function $`𝒫0`$ with $`𝒫(u)𝑑u=1`$, and the prefactor guarantees normalization. It follows that the moments of $`h`$ all scale with the mean height, $`\overline{h^n}=\overline{u^n}[\overline{h}(t)]^n`$ with $`\overline{u^n}`$ the $`n`$-th moment $`𝒫`$. In particular, the mean-square width $$W^2(t,L)=\text{var}(h)[\overline{h}(t)]^2,$$ (7) if Eq. (6) holds. On the other hand, we have that for times $`t<\tau `$ in the critical contact process, $$\overline{h}(t)=_0^t𝑑t^{}\rho (t^{})t^{1\theta },$$ (8) which immediately implies that $`W^2t^{2\beta _W}`$ with $`\beta _W=1\theta `$. In surface growth studies the crossover time is expected to scale as $`t_\times L^z`$; for a process in the DP universality class, we may write $`z=\nu _{||}/\nu _{}`$ (clearly $`t_\times `$ and $`\tau `$ should scale with the same exponent). Then the roughness exponent $`\alpha `$, defined via $`W^2(t,L)=W_{sat}^2(L)L^{2\alpha }`$ for $`tt_\times `$, is given by the scaling relation $`\alpha =\beta _Wz=(1\theta )\nu _{||}/\nu _{}`$. It is perhaps worth stressing that the expressions relating $`\alpha `$ and $`\beta _W`$ to DP exponents depend on the validity of the scaling hypothesis, Eq. (6), which should be tested. Inserting the known DP exponent values, scaling theory yields $`\alpha 1.3287`$, 0.97, 0.51, and zero for dimensions 1, 2, 3, and 4, respectively. In other words, the interface associated with the critical contact process is super-rough in one dimension, and asymptotically flat in $`dd_c=4`$, where $`\theta =1`$. The family of height-difference correlation functions $$G_q(r,t)\overline{|h(x,t)h(x+r,t)|^q},$$ (9) are also much studied in surface growth processes. Starting with a flat interface at $`t=0`$, we expect power-law growth, $`G_qr^{q\alpha _q}`$, for $`r<\xi (t)t^{1/z}`$, the time-dependent correlation length. If $`\alpha _q`$ depends on $`q`$ the interface is said to be multi-affine. For $`r>\xi `$, $`G_q`$ will saturate; in particular, $`G_2`$ should approach $`W^2(t,L)t^{2\beta _W}\xi ^{2\alpha }`$ for $`r\xi `$ and $`tt_\times `$. Since $`\xi `$ is the only length-scale relevant to correlations at short times (i.e., for $`t<t_\times `$, so that the system size $`L`$ does not come into play), it is reasonable to expect the scaling form $$G_2(r,t)=\xi ^{2\alpha }𝒢(r/\xi ),$$ (10) where the scaling function $`𝒢(x)x^{2\alpha _2}`$ for small $`x`$ and is constant for large $`x`$. The case $`\alpha _2=\alpha `$ is referred to as “conventional” scaling while $`\alpha _2<\alpha `$ is characterized as “anomalous” scaling . ## III Simulation Results We determined the square width $`W^2(t,L)`$ for rings of $`L=500`$, 1000, 2000, and 5000 sites, in samples of 2000, 1000, 1000, and 400 trials, respectively. The maximum time ranged from about $`1.6\times 10^5`$ for $`L=500`$ to $`8.8\times 10^6`$ for $`L=5000`$. All simulations were performed at the critical point, $`\lambda _c3.297848`$. In the contact process, interfacial properties can be studied over the full sample, or over a sample restricted to those trials that survive to time $`t`$. When a trial enters the absorbing state, $`W^2`$ naturally remains fixed for all subsequent times, and since all trials do eventually reach the absorbing state in the critical contact process, the interface width saturates for large $`t`$. If we restrict the sample to surviving trials, however, there is no saturation. It is important to note that the same scaling laws apply in either case. (The power-law growth regime, for example, corresponds to times $`<\tau `$, for which all trials still survive.) Fig. 1 shows a series of snapshots of the interface in a single trial with $`L=200`$, at intervals of 5000 time units. The progressive roughening of the interface, without evidence of saturation, is apparent. Fig. 2 is a scaling plot of the square width, averaged over all trials, i.e., $`W^2(L,t)/L^{2\alpha }`$ versus $`\stackrel{~}{t}t/L^z`$, using the exponents $`\alpha =1.328`$ and $`z=1.5808`$ expected for DP in 1+1 dimensions. There is a near-perfect data collapse for $`\stackrel{~}{t}>10^3`$. The power-law portion of the graph ($`10^3\stackrel{~}{t}0.05`$) yields the growth exponent $`\beta _W=0.839(1)`$, in good agreement with the value $`1\theta =0.8405`$ expected on the basis of the scaling argument. (Figures in parentheses denote statistical uncertainties.) Analysis of the saturation width yields $`W_{sat}^2(L)L^{2\alpha }`$ with $`\alpha =1.325(15)`$. In Fig. 3 we test the scaling assumption for the height probability distribution $`p(h)`$ by plotting $`t^{0.84}p(h)`$ versus $`h/t^{0.84}`$. (Recall that $`\overline{h}`$ is expected to grow $`t^{1\theta }=t^{0.84}`$.) In this system of $`L=1000`$ sites, there is a near-perfect data collapse, in accord with Eq. (6), for times between 500 and $`10^4`$. For $`t=2\times 10^4`$ we begin to note a departure from scaling, which becomes more pronounced at later times. Note that $`t=2\times 10^4`$ corresponds to $`\mathrm{ln}\stackrel{~}{t}=1`$, which is where $`W^2`$ begins to depart noticeably from a power law in Fig. 2. Analysis of $`p(h)`$ for $`L=5000`$ yields similar results. The form of the scaling function for a given reduced time $`\stackrel{~}{t}`$ is independent of the system size, as shown in Fig. 4, which compares height probability distributions for systems of 1000 and 5000 sites at times corresponding to the same $`\stackrel{~}{t}`$. The height-height correlation function $`G_2`$ in a system of 1000 sites is shown in Fig. 5, which is a double-logarithmic plot of $`G_2^{}G_2/t^{2\beta _W}`$ versus $`rr/t^{1/z}`$. There is a perfect data collapse for $`t`$ = 500,…,$`10^4`$ (reduced times $`10^2\stackrel{~}{t}0.2`$), with a power-law portion $`G_2r^{2\alpha _2}`$ with $`\alpha _20.625`$. At later times $`G_2^{}`$ does not collapse, principally because the square width has begun to saturate (it no longer grows $`t^{2\beta _W}`$). The value of $`\alpha _2`$ is insensitive to system size. For $`L=5000`$ we obtain $`\alpha _2=0.623(2)`$ for $`t=2\times 10^5`$, and 0.644 for $`t=5\times 10^5`$. Thus we may, with a high degree of confidence, adopt the estimate $`\alpha _2=0.63(3)`$, clearly much smaller than the roughness exponent $`\alpha =1.33`$ found in the analysis of the square width, indicating that this one-dimensional system exhibits anomalous scaling. Recently, López argued that anomalous surface roughening is associated with a diverging height gradient . In view of the growing spikes evident in the profiles shown in Fig. 1, this association seems very probable in the present instance. Indeed, the mean-square height gradient $$s(t)=\overline{(h)^2}.$$ (11) diverges as a power law (see Fig. 6). We find $`s(t)t^{2\kappa }`$, with $`\kappa =0.439(1)`$ for $`L=1000`$, $`0.4335(4)`$ for $`L=2000`$, and 0.4337(4) for $`L=5000`$; we adopt $`\kappa =0.4336(4)`$ as our best estimate. From López’ analysis, one expects $$\alpha _q=\alpha z\kappa .$$ (12) Inserting our result for $`\kappa `$ and the DP values $`\alpha =1.32867(14)`$ and $`z=1.5808(1)`$ in the r.h.s., we obtain $`\alpha _q=0.643(1)`$, in good agreement with the result found from analysis of the correlation function. We also studied the generalized height-height correlation function $`G_q`$, Eq. (9), for $`q`$ = 1/2, 1, 3/2, 2 and 3, in a system with $`L=1000`$, at $`t=10^4`$, and found that the functions $`[G_q]^{1/q}`$ are identical, and give $`\alpha _q0.62`$. We may therefore conclude that the DP interface is self-affine not multi-affine. In two dimensions we studied systems of up to 256$`\times `$256 sites at the critical point, $`\lambda _c1.6488`$. The curves for $`W^2`$ again show a good collapse (see Fig. 7), and the derived exponents are in very good agreement with the expected values. Specifically, we find $`\alpha =0.97(1)`$, $`\beta _W=0.550(5)`$, and $`z=1.765(10)`$. Scaling relations combined with known DP exponent values yield $`\alpha =0.970(5)`$ and $`\beta =0.549(2)`$, while current best estimates give $`z=\nu _{||}/\nu _{}=1.766(2)`$ for DP in 2+1 dimensions . For smaller system sizes we observe a transient in $`W^2`$ and all other measured quantities, which does not appear in one dimension. Anomalous roughening is also seen in two and three dimensions. Fig. 8 shows $`s(t)`$ growing $`t^{2\kappa }`$ with $`\kappa =0.33(1)`$. The height-height correlation function exhibits a good collapse, as shown in Fig. 9; the initial power-law growth yields $`\alpha _2=0.385(5)`$; thus the scaling relation Eq. (12) is well satisfied: $`\alpha _2\alpha +z\kappa =0.00(3)`$. We determined the interface exponents in three dimensions ($`\lambda _c1.3169`$, system size $`50^3`$ sites, maximum time 10<sup>4</sup>), though to somewhat lower precision, owing to the larger computational demand. The scaling relation yields $`\beta _W=1\delta =0.274(1)`$, while our simulation results for $`W^2`$ give $`\beta _W=0.27(1)`$. In three dimensions we find $`\alpha =0.51(1)`$, $`z=1.90(5)`$, and $`\kappa =0.22(1)`$. Together with Eq. (12), these yield $`\alpha _2=0.09(2)`$. Our results for critical exponents are collected in Table Interface Scaling in the Contact Process. ## IV Summary Defining an interface representation for the contact process by analogy with similar representations for sandpile models, we verified the expected scaling relation $`\beta _W=1\theta `$ in dimensions 1 - 3, and the scaling property of the height probability distribution in one dimension. The local roughness exponent, $`\alpha _2`$, is smaller than the global value, $`\alpha `$, indicating anomalous surface growth. This anomalous scaling is attended by a diverging local slope, $`s(t)=\overline{(h)^2}t^{2\kappa }`$. Our results for $`\kappa `$ are consistent with the scaling relation Eq. (12) derived by López. There is, on the other hand, no evidence of multi-affinity in this process. An interesting point is that the process continues to exhibit anomalous scaling for $`d=2`$ and 3, even though $`\alpha <1`$ in these cases. While it was pointed out some time ago that $`\alpha >1`$ implies anomalous growth , the latter appears to be an intrinsic feature of the contact process (and, by extension, of other models in the DP universality class) below $`d_c`$. Finally, we note that the anomalous roughening analysis introduces two new critical exponents, $`\alpha _2`$ and $`\kappa `$, and only one new scaling relation between them. We therefore have a new independent exponent, $`\kappa `$ for instance, that is not related to the standard DP exponents in any way. A very interesting theoretical task is that of computing $`\kappa `$ in an epsilon expansion around the upper critical dimension $`d_c=4`$. Our guess is that this new anomalous exponent is related to the renormalization of a composite operator not consider so far in the analysis of the Reggeon field theory, but this issue is beyond the scope of this paper, and will be studied elsewhere. Acknowledgements We are grateful to A. Allbens, J. Krug, J. M. López and J. G. Moreira for helpful comments. This work was supported in part by CNPq, by the European Network Contract ERBFMRXCT980183, and by the Ministerio de Educación, under project DGESEIC, PB97-0842. CAPES is acknowledged for support of computational facilities. electronic address: dickman@fisica.ufmg.br electronic address: mamunoz@onsager.ugr.es FIGURE CAPTIONS FIG. 1. Interface of the critical contact process in a system of 200 sites, shown at intervals of 5000 time units. FIG. 2. Scaled mean-square width versus reduced time for system sizes $`L=500`$, 1000, 2000, and 5000. FIG. 3. Scaling plot of the height probability distribution (unnormalized) for $`L=1000`$ at (from top to bottom on left-hand side) times 500, 1000, 2000, 5000, $`10^4`$, $`2\times 10^4`$, $`5\times 10^4`$, and $`10^5`$. FIG. 4. Scaling plot of the height probability distributions (unnormalized) for $`L=1000`$ (dashed lines) and $`L=5000`$ (solid lines) at reduced times (from top to bottom on left-hand side) $`\stackrel{~}{t}=0.142`$, 0.71, and 1.42. FIG. 5. Scaled height-height correlation function versus $`r/t^{1/z}`$ for $`L=1000`$. The topmost curve comprises collapsed data for times 500, 1000, 2000, 5000, and $`10^4`$; below it lie results for $`t=2\times 10^4`$ (solid curve), $`5\times 10^4`$ (dotted curve), and $`10^5`$ (dashed curve). FIG. 6. Growth of the mean-square gradient $`s(t)=\overline{(h)^2}`$ in systems with $`L=1000`$ (+), 2000 (dashed line), and 5000 (solid line). FIG. 7. Scaled mean-square width versus reduced time in two dimensions; system sizes $`L=32`$, 64, 128, and 256. FIG. 8. Growth of the mean-square gradient $`s(t)=\overline{(h)^2}`$ in the two-dimensional system with $`L=256`$. FIG. 9. Scaled height-height correlation function versus $`r/t^{1/z}`$ for $`L=256`$ in two dimensions. The curves (bottom to top) correspond to $`r=2`$, 4, 8, and 16.
warning/0003/cond-mat0003057.html
ar5iv
text
# Statistical Properties of Eigenstates in three-dimensional Mesoscopic Systems with Off-diagonal or Diagonal Disorder ## Abstract The statistics of eigenfunction amplitudes are studied in mesoscopic disordered electron systems of finite size. The exact eigenspectrum and eigenstates are obtained by solving numerically Anderson Hamiltonian on a three-dimensional lattice for different strengths of disorder introduced either in the potential on-site energy (“diagonal”) or in the hopping integral (“off-diagonal”). The samples are characterized by the exact zero-temperature conductance computed using real-space Green function technique and related Landauer-type formula. The comparison of eigenstate statistics in two models of disorder shows sample-specific details which are not fully taken into account by the conductance, shape of the sample and dimensionality. The wave function amplitude distributions for the states belonging to different transport regimes within the same model are contrasted with each other as well as with universal predictions of random matrix theory valid in the infinite conductance limit. The disorder induced localization-delocalization (LD) transition in solids has been one of the most vigorously pursued problems in condensed matter physics since the seminal work of Anderson. In thermodynamic limit, strong enough disorder generates a zero-temperature critical point in $`d>2`$ dimensions as a result of quantum interference effects (in $`d2`$ even weakly disordered metal turns into an Anderson insulator for sufficiently large sample size). Thus, research in the “pre-mesoscopic” era was mostly directed toward the viewpoint provided by the theory of critical phenomena. The advent of mesoscopic quantum physics has unearthed large fluctuations, induced by quantum coherence and randomness of disorder, of various physical quantities (e.g., conductance, local density of states, current relaxation times, etc.), even well into the delocalized phase. Thus, complete understanding of the LD transition requires to examine full distribution functions of relevant quantities. Especially interesting are the deviations of their asymptotic tails, a putative signature of incipient localization, from the (usually) Gaussian distributions expected in the limit of infinite dimensionless conductance $`g=G/G_Q`$ ($`G_Q=2e^2/h`$ is the conductance quantum). This paper presents the study of such type—numerical computation of the statistics of eigenfunction amplitudes in finite-size three-dimensional (3D) nanoscale (composed of $`1000`$ atoms) mesoscopic disordered conductors. The 3D conductors are often “neglected” in favor of the more popular and tractable playgrounds—two-dimensional systems (2D), where one can study states resembling 3D critical wave functions in a wide range of systems sizes and disorder strengths, or quasi one-dimensional systems where analytical techniques can handle even non-perturbative phenomena (like the ones at small $`g`$). For example, in $`d=2+ϵ`$ dimensions LD transition occurs at weak disorder (weak-coupling regime of the corresponding field-theoretical description ), while in $`d3`$ small parameter needed for analytical treatment is lacking. In 3D systems critical eigenfunctions, exhibiting multifractal scaling, are expected only at the mobility edge $`E_c`$ which separates extended and localized states inside the energy band. The essential physics of disordered conductors is captured by studying the quantum dynamics of a non-interacting (quasi)particle in a random potential. This problem is classically non-integrable, thereby exhibiting quantum chaos. The concepts unifying disordered electron physics with the standard ‘clean’ (i.e., without stochastic disorder) examples of quantum chaos come from statistical approaches to the properties of energy spectrum and corresponding eigenstates, which cannot be computed analytically. While level statistics of disordered systems have been explored to a great extent, investigation of the statistics of eigenfunctions has been initiated only recently. These studies are not only revealing peculiar spectral properties of random Hamiltonians, but are relevant for the thorough understanding of various unusual features of quantum transport in diffusive metallic conductors (including the ones which are proximity coupled to a superconductor ). The standard examples are long-time tails in the relaxation of current or log-normal tails (in $`d=2+ϵ`$) of the distribution function of mesoscopic conductances. Distribution of eigenfunction amplitudes is found to be relevant for tunneling measurements on quantum dots probing the coupling to external leads, which depends sensitively on the local properties of wave functions. Experiments which are the closest to directly delving into the microscopic structure of quantum chaotic or disordered wave functions exploit the correspondence between the Schrödinger and Maxwell equations in microwave cavities. The study of fluctuations and correlations of eigenfunction amplitudes in mesoscopic systems has led to a concept of the so-called “pre-localized” states. In 3D delocalized phase, this notion refers to the states which have sharp amplitude peaks on the top of an extended background. These kind of states appear even in the diffusive, $`\mathrm{}L<\xi `$, metallic ($`g1`$) regime, but are anomalously rare in such samples (here $`\mathrm{}`$ is the elastic mean free path and $`\xi `$ is the localization length). In order to get “experimental” feeling for the structure of states with unusually high amplitude spikes, an example is given in Fig. 1; this state is found in a special realization of quenched disorder (out of many randomly generated impurity configurations) inside the sample characterized by large average conductance. Thus, pre-localized states are putative precursors of LD transition and determine asymptotics of some of the distribution functions studied in open or closed mesoscopic systems. In $`d2`$, where all states are considered to be localized, pre-localized states have anomalously short localization radius when compared to “ordinary” localized states in low-dimensional systems. They underlie the multifractal scaling in weakly localized ($`g1`$) 2D conductors of size $`L`$ smaller than exponentially large $`\xi `$ (which plays the role of a phase transition correlation length $`\xi _c=\xi `$ in $`d2`$). In 3D, the correlation length $`\xi _c`$ (defined as the size of a hypercube for which $`g(\xi _c)=𝒪(1)`$) is always microscopic ($`\xi _c\lambda _F`$) in good metallic samples, and no multifractal scaling is expected. The appearance of small regions inside disordered solids where eigenstates can have large amplitudes seems to be a “strongly pronounced” analog of the phenomenon of scarring (anomalous enhancement or suppression of quantum chaotic wave function intensity on the unstable periodic orbits in the corresponding classical system) introduced in the guise of generic quantum chaos. In general, the study of properties of wave functions on a scale smaller than $`\xi `$ should probe quantum effects causing evolution of extended into localized states upon approaching the LD critical point. In the marginal two-dimensional case, the divergent (in the limit $`L\mathrm{}`$) weak localization (WL) correction to the semiclassical Boltzmann conductivity provides an explanation of localization in terms of the interference between two amplitudes to return to initial point along the same classical path in the opposite directions. This simple quantum interference effect leads to a coherent backscattering (i.e., suppression of conductivity) in a time-reversal invariant systems without spin-orbit interaction. However, in 3D systems WL correction is not “strong” enough to provide a full microscopic picture of complicated quantum interference processes which are responsible for LD transition, and facilitate the expansion of quantum intuition. The paper presents the statistics of eigenfunction intensities $`|\mathrm{\Psi }_\alpha (𝐫)|^2`$ in isolated 3D mesoscopic conductors characterized by two different types of microscopic disorder. Numerical methods employed here make it possible to treat phenomena in both semiclassical (described by Bloch-Boltzmann formalism and perturbative quantum corrections ) and fully quantum transport regime (dominated by non-perturbative effects, where semiclassical concepts, like $`\mathrm{}`$, loose their meaning ), as well as in the crossover realm. Since mesoscopic physics has provided efficient techniques for “measuring” exactly the transport properties of finite-size samples on the computer, this study connects the eigenstates statistics of a closed sample to its zero-temperature conductance. The statistical properties of eigenstates are described by the disorder-averaged distribution function $$f(t)=\frac{1}{\rho (E)N}\underset{𝐫,\alpha }{}\delta (t|\mathrm{\Psi }_\alpha (𝐫)|^2V)\delta (EE_\alpha ),$$ (1) on $`N`$ discrete points $`𝐫`$ inside a sample of volume $`V`$. Here $`\rho (E)=_\alpha \delta (EE_\alpha )`$ is the mean level density at energy $`E`$. Averaging over disorder is denoted by $`\mathrm{}`$. Normalization of eigenstates gives $`\overline{t}=𝑑ttf(t)=1`$. A finite-size disordered sample is modeled by a tight-binding Hamiltonian (TBH) with nearest neighbor hopping integral $`t_{\mathrm{𝐦𝐧}}`$ $$\widehat{H}=\underset{𝐦}{}\epsilon _𝐦|𝐦𝐦|+\underset{𝐦,𝐧}{}t_{\mathrm{𝐦𝐧}}|𝐦𝐧|,$$ (2) on a simple cubic lattice $`16\times 16\times 16`$ of lattice constant $`a`$. Each site $`𝐦`$ contains a single $`s`$-orbital $`𝐫|𝐦=\psi (𝐫𝐦)`$. Periodic boundary conditions are chosen in all directions. In a random hopping (RH) model the disorder is introduced by taking the off-diagonal matrix elements to be a uniformly distributed random variable, $`12W_{\text{RH}}<t_{\mathrm{𝐦𝐧}}<1`$, while diagonal elements are zero $`\epsilon _𝐦=0`$. The strength of the disorder is measured by $`W_{\text{RH}}`$. The other system studied is described by a diagonally disordered (DD) Anderson model with potential energy $`\epsilon _𝐦`$ on site $`𝐦`$ drawn from the uniform distribution, $`W_{\text{DD}}/2<\epsilon _𝐦<W_{\text{DD}}/2`$, and $`t_{\mathrm{𝐦𝐧}}=1`$ is the unit of energy. The Hamiltonian (2) is a real symmetric matrix because time-reversal symmetry is assumed. The results for $`f(t)`$ in the samples described by the RH and DD Anderson models are shown on Figs. 23 and Fig. 4, respectively. Although some of the samples are characterized by similar values of conductance, the eigenstates in the two models show different statistical behavior. In what follows the meaning of these findings is explained in the context of statistical approach to quantum systems with non-integrable classical dynamics. In particular, the results are compared to the universal predictions of random matrix theory (RMT). In the statistical approach of RMT, the Hamiltonian of a quantum chaotic system is replaced by a random matrix drawn from an ensemble defined by the symmetries under time-reversal and spin-rotation. This leads to the Wigner-Dyson (WD) statistics for eigenvalues and Porter-Thomas (PT) distribution for eigenfunction intensities. For the Gaussian orthogonal ensemble (GOE), relevant for studies of time-reversal-invariant Hamiltonians like (2), the PT distribution is given by $$f_{\text{PT}}(t)=\frac{1}{\sqrt{2\pi t}}\mathrm{exp}(t/2).$$ (3) The function $`f_{\text{PT}}(t)`$ is plotted as a reference in Figs. 23 and Fig. 4. The predictions of RMT are universal, depending only on the symmetry properties of the relevant ensemble. They apply to the statistics of real disordered systems in the limit $`g\mathrm{}`$ ($`g=\pi E_{\text{Th}}/\mathrm{\Delta }`$, where $`\mathrm{\Delta }=1/\rho (E)`$ is the mean energy level spacing and $`E_{\text{Th}}=\mathrm{}𝒟/L^2`$ is the Thouless energy, set by the classical diffusion across a sample of size $`L`$ with diffusion constant $`𝒟`$). The spectral correlations in RMT are determined by logarithmic level repulsion which is independent of true dynamics. All sample-specific details are absorbed into the mean level spacing $`\mathrm{\Delta }`$. Also, the level correlations are independent of the eigenstate correlations. The RMT answer (3) for the distribution function (1) was derived by Porter and Thomas by assuming that the coordinate-representation eigenstate $`𝐫|\mathrm{\Psi }_\alpha `$ in a disordered (or classically chaotic system) is a Gaussian random variable. The behavior of $`\mathrm{\Psi }_\alpha (𝐫)`$, even within the framework of RMT, is simple only in the systems with unbroken or completely broken time-reversal symmetry \[the only difference between the two limiting ensembles is the functional form of $`f(t)`$\]. Thus, RMT implies statistical equivalence of eigenstates which equally test the random potential all over the sample—typical wave function has more or less uniform amplitude $`1/\sqrt{V}`$, up to inevitable Gaussian fluctuations. Microscopic theory brings corrections to the RMT results in the case of samples with finite $`g`$. In the finite-size systems level statistics follow RMT predictions in the ergodic regime, i.e., on the energy separation scale smaller than $`E_{\text{Th}}`$. Non-universal corrections to the spectral statistics or eigenfunction statistics (which describe the long-range correlations of wave functions) depend on dimensionality, shape of the sample, and conductance $`g`$. These deviations from RMT predictions grow with increasing disorder (i.e., lowering of $`g`$). At the LD transition wave functions acquire multifractal properties, while the critical level statistics become scale-independent. For strong disorder or, at fixed disorder, for energies $`|E|`$ above the mobility edge $`|E_c|`$, wave functions are exponentially localized. The simple (and usually invoked) picture is that of a wave function which decays as $`\mathrm{\Psi }(r)=p(r)\mathrm{exp}(r/\xi )`$ from its maximum centered at some point inside the sample of size $`L>\xi `$. Here $`p(r)`$ is a random function and approximately spherical symmetry of decay is assumed. Since two states close in energy are localized at different points in space, there is almost no overlap between them. Therefore, the levels become uncorrelated and obey Poisson statistics. If $`p(r)=c`$ is simplified to a normalization constant, the distribution function of intensities is given by $`f_\xi (t)`$ $`=`$ $`{\displaystyle \frac{4\pi }{V}}{\displaystyle \underset{0}{\overset{L/2}{}}}𝑑rr^2\delta (t|\mathrm{\Psi }(r)|^2V)={\displaystyle \frac{\pi \xi ^3}{4V}}{\displaystyle \frac{\mathrm{ln}^2(c^2V/t)}{t}},`$ (4) $`c^2`$ $`=`$ $`{\displaystyle \frac{2}{\pi \xi ^3}}\left[1\left(1+{\displaystyle \frac{L}{\xi }}+{\displaystyle \frac{L^2}{2\xi ^2}}\right)e^{L/\xi }\right]^1,`$ (5) where a spherically symmetric sample of radius $`L/2`$ is assumed. In the localized phase $`\xi L`$, $`f_\xi (t)`$ is expected to be insensitive to the assumed shape of the sample, and is determined by the ratio of these two relevant length scales. The distribution function $`f(t)`$ is equivalently given in term of its moments $`b_q=𝑑tt^qf(t)`$ (this statement is not rigorous since examples of different distribution functions which possess exactly the same sets of moments are encountered in various statistical problems ). For GOE, the PT distribution (3) has moments $`b_q^{\text{PT}}=2^qV^{q+1}\mathrm{\Gamma }(q+1/2)/\mathrm{\Gamma }(1/2)`$. They are related to the moments $`I_\alpha (q)=𝑑𝐫|\mathrm{\Psi }_\alpha (𝐫)|^{2q}`$ of the wave function intensity. In the universal regime $`g\mathrm{}`$ wave functions cover the whole volume with only short-range correlations (on the scale $`|𝐫_1𝐫_2|\mathrm{}`$) persisting between $`\mathrm{\Psi }_\alpha (𝐫_1)`$ and $`\mathrm{\Psi }_\alpha (𝐫_2)`$. This means that integration in the definition of $`I_\alpha (q)`$ provides self-averaging, and $`I_\alpha (q)`$ does not fluctuate in the universal limit, i.e., $`I_\alpha (q)=b_q^{\text{PT}}`$. On the other hand, at finite $`g`$ spatial correlations of wave function amplitudes at distances comparable to the system size are non-negligible. Therefore, $`I_\alpha (q)`$ fluctuates from state to state and from sample to sample. Although these long-range spatial correlations necessitate to study the full distribution function of $`I_\alpha (q)`$, for the subsequent analysis in this study it is enough to use an ensemble average of $`I_\alpha (q)`$, i.e., following Wegner $$I(q)=\mathrm{\Delta }\underset{𝐫,\alpha }{}|\mathrm{\Psi }_\alpha (𝐫)|^{2q}\delta (EE_\alpha ).$$ (6) The moment $`I_\alpha (2)`$ is usually called inverse participation ratio (IPR). It is a one-number measure of the degree of localization (i.e., it measures the portion of space where the amplitude of the wave function differs markedly from zero). This becomes obvious from the scaling properties of the average moments $`I(q)`$ with respect to the system size $$I(q)\{\begin{array}{cc}L^{d(q1)}\text{ metal},\hfill & \\ L^0\text{insulator},\hfill & \\ L^{d^{}(q)(q1)}\text{critical}.\hfill & \end{array}$$ (7) Here $`d^{}(q)<d`$ is the fractal dimension. Its dependence on $`q`$ is the hallmark of multifractality of wave functions. The multifractal wave functions are delocalized, but extremely inhomogeneous occupying only an infinitesimal fraction of the sample volume in thermodynamic limit. The IPR is affected by mesoscopic fluctuations which scale in metallic samples as $`\delta I_\alpha (2)1/g^2L^{42d}`$. In the critical region ($`g1`$) fluctuations are of the same order as the average value, which is then not enough to characterize the critical eigenstates (even though multifractal wave function extend throughout the whole sample, their IPR is not self-averaging ). I use $`I(2)`$ as a rough guide in selecting eigenstates with different properties in the delocalized phase (Fig. 5). The second parameter used in the selection procedure is the conductance $`g(E_F)`$ computed for a band filled up to the Fermi energy $`E_F`$ equal to the state eigenenergy (see Fig. 6). The conductance as a function of band filling allows one to delineate delocalized from localized phase as well as to narrow down the critical region around LD transition point (which is defined by $`g(E_c)1`$). Upon inspection of these two parameters, a small window is placed around chosen energy, and $`f(t)`$ is computed for all eigenenergies whose eigenvalues fall inside the window. This provides more detailed information on the structure of eigenstates than is encoded in IPR. The average IPR for both RH and DD Anderson model is shown in Fig. 5. The models with random hopping have been around for a long time, but have attracted considerable attention only recently, inasmuch as they show a disorder induced quantum critical point in less than three dimensions, where delocalization occurs in the band center. Furthermore, several models used for strongly correlated electron systems share similar mathematical structure with the random hopping 2D models formulated in the framework of (non-interacting) disorder electron physics. Real solids which could be described by TBH (2) with off-diagonal disorder include doped semiconductors, such as P-doped Si, where hopping integrals $`t_{\mathrm{𝐦𝐧}}`$ vary exponentially with the distances between the orbitals they connect, while diagonal on-site energies $`\epsilon _𝐦`$ are nearly constant. The behavior of low-dimensional RH Anderson model goes against the standard mantra of the scaling theory of localization for non-interacting systems that all electron states are localized in $`d2`$ for arbitrarily weak disorder. The possibility of delocalization transition in one dimension goes back to the work of Dyson on glasses. Also, the scaling theory for quantum wires with off-diagonal disorder requires two parameters which depend on the microscopic model, thus breaking the celebrated universality in disordered electron problems. In 3D case explored here, the states in the band center are less extended than other delocalized states inside the band (Fig. 5). However, the off-diagonal disorder is not strong enough to localize all states in the band, in contrast to the usual case of diagonal disorder where whole band becomes localized for $`W_{\text{DD}}^c16.5`$. The mobility edge for the strongest RH disorder $`W_{\text{RH}}=1`$, as well as for DD models, is found by looking at an exact zero-temperature static conductance. This quantity (which is a Fermi surface property) is computed from the Landauer-type formula $$g(E_F)=\text{Tr}[𝐭(E_F)𝐭^{}(E_F)],$$ (8) where transmission matrix $`𝐭(E_F)`$ is expressed in terms of the real-space (lattice) Green functions for the sample attached to two clean semi-infinite leads. To study the conductance in the whole band of the DD model, $`t_{\mathrm{𝐦𝐧}}=1.5`$ is used for the hopping integral in the leads. This mesoscopic computational technique “opens” the sample, thereby smearing the discrete levels of an initially isolated system. Thus, the spectrum of sample+leads=infinite system becomes continues and conductance can be calculated at any $`E_F`$ inside the band. However, the computed conductance, for not too small disorder or coupling to the leads (which are of the same transverse width as the sample ) is practically equal to the “intrinsic” conductance $`g=\pi E_{\mathrm{Th}}/\mathrm{\Delta }`$ determined by the spectral properties of a closed sample. The conductance and density of states (DoS) $$N(E)=2\frac{\rho (E)}{V},$$ (9) are plotted in Fig. 6 for the samples whose eigenstates are investigated (the factor of two is for spin degeneracy). The DoS is obtained from the histogram of the number of eigenvalues which fall into equally spaced energy bins along the band. The conductance and DoS of the RH model have a peak at $`E=0`$, which becomes a logarithmic singularity in the limit of infinite system size. To get an insight into the “general weakness” of the off-diagonal disorder, Fig. 7 plots $`g(E_F)`$ and $`N(E)`$ for the low $`W_{\text{RH}}=0.25`$. In this case $`N(E)`$ still resembles the DoS of a clean system, even after ensemble averaging. On the other hand, the conductance is a smooth function of energy since discrete levels of an isolated sample are broadened by the coupling to leads. The same is true for the DoS computed from the imaginary part of the Green function for an open system. The mobility edge is absent at low RH disorder ($`W_{\mathrm{RH}}=0.25`$ and $`W_{\mathrm{RH}}=0.375`$) for system sizes $`L16a`$. This means that localization length $`\xi `$ is greater than $`16a`$ for all energies inside the band of these systems. For other samples on Fig. 6 the mobility edge appears inside the band. This is clearly shown for $`W_{\mathrm{RH}}=1`$ case where band edge $`E_b`$ ($`N(E_b)=0`$) differs from $`E_c`$. The mobility edge is located at the minimum energy $`|E_c|`$ for which $`g(E_c)`$ is still different from zero. The conductance of finite samples is always finite, although exponentially small at $`|E_F|>|E_c|`$. The approximate values of $`|E_c|`$ listed in Fig. 6 are such that conductance satisfies: $`g(E_F)<0.1`$, for $`|E_F|>|E_c|`$; typically $`g(E_c)(0.2,0.5)`$ is obtained, like in the recent detailed studies of conductance properties at $`E_c`$. Thus found $`E_c`$ is virtually equal to the true mobility edge, which is properly defined only in thermodynamic limit (and usually obtained from some numerical finite-size scaling procedure ). Namely, the position of mobility edge extracted in this way will not change when going to larger system sizes if $`\xi <L`$ for all energies $`|E|>|E_c|`$. The distribution $`f(t)`$ of eigenfunction intensities has been studied analytically for diffusive conductors close to the universal RMT limit (where conductance is large and localization effects are small) in Refs. using the supermatrix $`\sigma `$-model (NLSM), or by means of a “direct optimal fluctuations method” of Ref. . Numerical studies of statistics of eigenstates in DD Anderson model were conducted in 2D for all disorder strengths, while in 3D the focus has been on the states appearing in the semiclassical transport regime where comparison with analytical predictions (parameterized by semiclassical quantities, like $`k_F\mathrm{}`$) can be made in the regions of small and large deviations of $`f(t)`$ from PT distribution. Here I show how $`f(t)`$ evolves with the strength of (different types of) disorder in 3D samples, where genuine LD transition occurs in the strong coupling regime of the corresponding field-theoretical formulation. In the weakly disordered ($`k_F\mathrm{}1`$) metallic ($`g1`$) conductors pre-localized states are extremely rare. For example, in an ensemble of 20 000 samples ($`W_{\mathrm{DD}}=4`$), whose typical transport properties are well-described by semiclassical theories, only four states would show up in the band center which exhibit similar amplitude splashes like the one in Fig. 1. Thus, to get a far tail (where deviations from PT distribution are large) of $`f(t)`$ in such conductors one has to search through enormous ensemble and locate special configurations of a random potential. The maximum wave function amplitudes which can be observed in this pursuit are, plausibly, determined by the strength of disorder, i.e., conductance $`g`$. It is, however, interesting that tails at small but finite $`g`$ in the strong disorder (like $`W_{\mathrm{DD}}=10`$) can be longer than the tails of states in the localized phase, where $`g`$ is vanishingly small, for $`|E|>|E_c|`$ at some smaller fixed disorder (like in the case of $`W_{\mathrm{DD}}=6`$; compare the two panels in Fig. 4 using respective conductances from Fig. 6). For strong enough disorder long tails of $`f(t)`$ are found, even by investigating small ensembles of disordered conductors (as shown below), since the frequency of appearance of pre-localized states is greatly enhanced (while system is still on the delocalized side of LD transition). The complete eigenproblem of a single particle random Hamiltonian is solved exactly by numerical diagonalization. Then, $`f(t)`$ is computed as a histogram of intensities for the chosen eigenstates in: delocalized ($`|E|<|E_c|`$), localized phase ($`|E|>|E_c|`$), and critical region around the mobility edge $`|E_c|`$. The two delta functions in Eq. (1) are approximated by a box function $`\overline{\delta }(x)`$. The width of $`\overline{\delta }(EE_\alpha )`$ is small enough at a specific energy that $`\rho (E)`$ is constant inside that interval. For each sample, 5–10 states are picked by the energy bin, which effectively provides additional averaging over the disorder (according to ergodicity in RMT). The amplitudes of wave functions are sorted in the bins defined by $`\overline{\delta }(t|\mathrm{\Psi }_\alpha (𝐫)|^2V)`$ whose width is constant on a logarithmic scale. The function $`f(t)`$ is computed at all points inside the sample, i.e., $`N=16^3`$ in Eq. (1). The evolution of $`f(t)`$, when sweeping the band through the interesting states, is plotted in Figs. 23 for the RH disordered sample. Since pre-localized states generate slow decay of $`f(t)`$ at high wave function intensities (where PT distribution is negligible), this region is enlarged on Fig. 3. This is obvious from the pre-localized example in Fig. 1 where state with large amplitude spikes, highly unlikely in the framework of RMT, was found in a very good metal. The same is trivially true for the localized states which determine extremely long tails of $`f_\xi (t)`$. Thus, the long asymptotic tails of $`f(t)`$, appreciably deviating from PT distribution, are signaling the onset of localization. It is interesting that states in the band center of RH model, which define the largest zero-temperature conductance $`[g(E_F=0)10.2`$, Var $`g(E_F=0)0.63]`$, are in fact mostly pre-localized. Moreover, both the frequency of their appearance and high amplitude splashes resemble the situation at criticality (NLSM-type calculation shows that 3D wave functions, sufficiently close to the band center, are always extended for any disorder strength). It might be conjectured that these pre-localized states would generate multifractal scaling of IPR in the band center. This result, together with the DoS and conductance from Figs. 67, shows that phenomena in the band center of 3D conductors with off-diagonal disorder are as intriguing as their much studied counterparts in low-dimensional systems. The origin of these phenomena can be traced back to a special sublattice, or “chiral”, symmetry obeyed by TBH (2) with random hopping and constant on-site energy (leading to an eigenspectrum which for $`E_\alpha `$ contains $`E_\alpha `$, with a special role played by $`E=0`$). In the $`W_{\text{RH}}=0.25`$ and $`W_{\text{RH}}=0.375`$ cases all states are extended. Their $`f(t)`$ looks similar to the distribution function for the delocalized states at $`E=1.5`$ in the sample characterized by $`W_{\text{RH}}=1`$. The distribution function $`f_\xi (t)`$ defined in Eq. (4) fits reasonably well the numerical $`f(t)`$ generated by the states around $`E=2.75`$, where $`\xi 1.2a`$ is extracted for the localization length. Thus, one can measure approximately $`\xi `$ in this way even though the structure of localized eigenstates can be more complicated than the simple radially symmetric exponential decay used to derive $`f_\xi (t)`$ \[e.g., in the case of DD disorder $`\xi 1.3a`$ is obtained for examined localized states in Fig. 4 for $`W_{\mathrm{DD}}=10`$ ensemble, while corresponding states in $`W_{\mathrm{DD}}=6`$ ensemble seem to be too close to the mobility edge to follow $`f_\xi (t)`$\]. The same statistical analysis is performed for the eigenstates of DD Anderson model—a “standard model” in the localization theory. Figure 4 plots $`f(t)`$ at specific energies $`E_i`$ in samples characterized by different conductances $`g(E_F=E_i)`$. The conductance $`g(E=0)`$ of TBH with $`W_{\text{DD}}=6`$ is numerically close to the conductance of RH disordered samples with $`W_{\text{RH}}=1`$. Nevertheless, comparison of the corresponding distribution functions reveals disorder-dependent features which are beyond universal corrections (i.e., independent of the details of random potential) accounted by the properties of a classical diffusion operator (the spectrum of $`𝒟^2`$, with appropriate boundary conditions, depends on $`g`$, shape of the sample and dimensionality; note that conductors at $`W6`$ with half-filled band lie on the boundary of applicability of such semiclassical concepts ). To get the far tail of the eigenstate statistics in the band center a much larger statistics is needed than used here. This then makes the observed $`f(t)`$ in a special case of the band center of $`W_{\text{RH}}=1`$ disordered conductor even more spectacular because of very large amplitudes found in the small ensemble of conductors. Thus, the strong dependence on the microscopic details of random potential demonstrated in this study is somewhat different from the disorder-specific short length scale (“ballistic” ) contributions to the standard picture of diffusive NLSM. Namely, here it seems that special features of off-diagonal disorder in the band center generate completely different functional form of the far tail, and not just some disorder-specific values of the parameters in the exp-log-cube asymptotics. . In both models, all computed $`f(t)`$ intersect PT distribution (from below) around $`6t10`$, and then develop tails far above PT values. The length of the tails is defined by the largest amplitude exhibited in the pre-localized state (like that in Fig. 1). For strong DD, $`W_{\text{DD}}=10`$, the conductance $`g(E_F)`$ is smaller than $`3.5`$. In this regime transport becomes “intrinsically diffusive”, but one can still extract resistivity from the approximate Ohmic scaling of disorder-averaged resistance (for those fillings where $`g(E_F)>2`$). However, the close proximity to the critical region $`g1`$ induces long tails of $`f(t)`$ at all energies throughout the band—a sign of increased frequency of appearance of highly inhomogeneous states. This provides an insight into the microscopic structure of eigenstates which carry the current in a non-semiclassical transport regime (characterized by the lack of simple intuitive concepts, like mean free path $`\mathrm{}`$, since unwarranted use of the Boltzmann theory would give $`\mathrm{}<a`$ in this transport regime although the sample is still far away from the LD transition). In conclusion, the statistics of eigenstates in 3D mesoscopic conductors, modeled by the tight-binding Hamiltonian on the finite-size simple cubic lattice, have been studied. The disorder is introduced either in the potential energy (diagonal) or in the hopping integrals (off-diagonal). Also calculated are the average inverse participation ratio of eigenfunctions and the exact zero-temperature conductance as a function of Fermi energy. This comprehensive set of parameters makes it possible to compare the eigenstates in nanoscale samples with different types of disorder, but characterized by similar values of conductance. Disorder-specific details, which are not parameterized by the conductance alone, are found. This is in spite of the fact that dimensionality, shape of the sample, and conductance are expected to determine the finite-size corrections to the universal predictions of random matrix theory, at least in the samples which are in the semiclassical transport regime. The appearance of states with large amplitude spikes on the of top of RMT like background is clearly demonstrated even in good metals. At criticality, the proliferation of such “pre-localized” states is directly related to the extensively studied multifractal scaling of IPR. However, even in the delocalized phase with good metallic properties ($`g1`$), where the correlation length $`\xi _c\lambda _F`$ defined by the sample conductance $`g(\xi _c)1`$ is microscopic ($`L<\xi _c`$ would naturally account for the multifractal scaling, like in 2D), pre-localized state with unexpectedly high amplitudes are found in the band center of random hopping disordered systems. They are inhomogeneous enough to generate extremely long tails of the distribution of eigenfunction amplitudes, akin to the ones observed at criticality. Inspiring discussions with V. Z. Cerovski are acknowledged. Valuable guidance have been provided by P. B. Allen. The important improvements of the initial cond-mat preprint have resulted from criticism provided by B. Shapiro and I. E. Smolyarenko. This work was supported in part by NSF grant no. DMR 9725037. $``$ Present address: Department of Physics, Georgetown University, Washington, DC 20057-0995.
warning/0003/quant-ph0003004.html
ar5iv
text
# Simple Proof of Security of the BB84 Quantum Key Distribution Protocol ## Abstract We prove that the 1984 protocol of Bennett and Brassard (BB84) for quantum key distribution is secure. We first give a key distribution protocol based on entanglement purification, which can be proven secure using methods from Lo and Chau’s proof of security for a similar protocol. We then show that the security of this protocol implies the security of BB84. The entanglement-purification based protocol uses Calderbank-Shor-Steane (CSS) codes, and properties of these codes are used to remove the use of quantum computation from the Lo-Chau protocol. Quantum cryptography differs from conventional cryptography in that the data are kept secret by the properties of quantum mechanics, rather than the conjectured difficulty of computing certain functions. The first quantum key distribution protocol, proposed in 1984 , is called BB84 after its inventors (C. H. Bennett and G. Brassard). In this protocol, the participants (Alice and Bob) wish to agree on a secret key about which no eavesdropper (Eve) can obtain significant information. Alice sends each bit of the secret key in one of a set of conjugate bases which Eve does not know, and this key is protected by the impossibility of measuring the state of a quantum system simultaneously in two conjugate bases. The original papers proposing quantum key distribution proved it secure against certain attacks, including those feasible using current experimental techniques. However, for many years, it was not rigorously proven secure against an adversary able to perform any physical operation permitted by quantum mechanics. Recently, three proofs of the security of quantum key distribution protocols have been discovered; however, none is entirely satisfactory. One proof , although easy to understand, has the drawback that the protocol requires a quantum computer. The other two prove the security of a protocol based on BB84, and so are applicable to near-practical settings. However, both proofs are quite complicated. We give a simpler proof by relating the security of BB84 to entanglement purification protocols and quantum error correcting codes . This new proof also may illuminate some properties of previous proofs , and thus give insight into them. For example, it elucidates why the rates obtainable from these proofs are related to rates for CSS codes. The proof was in fact inspired by the observation that CSS codes are hidden in the inner workings of the proof given in . We first review CSS codes and associated entanglement purification protocols. Quantum error-correcting codes are subspaces of the Hilbert space $`^{2^n}`$ which are protected from errors in a small number of these qubits, so that any such error can be measured and subsequently corrected without disturbing the encoded state. A quantum CSS code $`Q`$ on $`n`$ qubits comes from two binary codes on $`n`$ bits, $`C_1`$ and $`C_2`$, one contained in the other: $`\{0\}C_2C_1𝐅_2^n,`$ where $`𝐅_2^n`$ is the binary vector space on $`n`$ bits . A set of basis states (which we call codewords) for the CSS code subspace can be obtained from vectors $`vC_1`$ as follows: $$v\frac{1}{|C_2|^{1/2}}\underset{wC_2}{}|v+w.$$ (1) If $`v_1v_2C_2`$, then the codewords corresponding to $`v_1`$ and $`v_2`$ are the same. Hence these codewords correspond to cosets of $`C_2`$ in $`C_1`$, and this code protects a Hilbert space of dimension $`2^{dimC_1dimC_2}`$. The above quantum code is equivalent to the dual code $`Q^{}`$ obtained from the two binary codes $`\{0\}C_1^{}C_2^{}𝐅_2^n.`$ This equivalence can be demonstrated by applying the Hadamard transform $`H={\displaystyle \frac{1}{\sqrt{2}}}\left(\begin{array}{cc}\hfill 1& \hfill 1\\ \hfill 1& \hfill 1\end{array}\right)`$ to each encoding qubit. This transformation interchanges the bases $`|\mathrm{\hspace{0.17em}0}`$, $`|\mathrm{\hspace{0.17em}1}`$ and $`|+`$, $`|`$, where $`|+=\frac{1}{\sqrt{2}}(|\mathrm{\hspace{0.17em}0}+|\mathrm{\hspace{0.17em}1})`$ and $`|=\frac{1}{\sqrt{2}}(|\mathrm{\hspace{0.17em}0}|\mathrm{\hspace{0.17em}1})`$. It also interchanges the two subspaces corresponding to the codes $`Q`$ and $`Q^{}`$, although the codewords (given by Eq. 1) of $`Q`$ and $`Q^{}`$ are not likewise interchanged. We now make a brief technical detour to define some terms. The three Pauli matrices are: $`\sigma _x=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right),\sigma _y=\left(\begin{array}{cc}\hfill 0& \hfill i\\ \hfill i& \hfill 0\end{array}\right),\sigma _z=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right).`$ The matrix $`\sigma _x`$ applies a bit flip error to a qubit, while $`\sigma _z`$ applies a phase flip error. We denote the Pauli matrix $`\sigma _a`$ acting on the $`k`$’th bit of the CSS code by $`\sigma _{a(k)}`$ for $`a\{x,y,z\}`$. For a binary vector $`s`$, we let $`\sigma _a^{[s]}=\sigma _{a(1)}^{s_1}\sigma _{a(2)}^{s_2}\sigma _{a(3)}^{s_3}\mathrm{}\sigma _{a(n)}^{s_n}`$ where $`\sigma _a^0`$ is the identity matrix and $`s_i`$ is the $`i`$’th bit of $`s`$. The matrices $`\sigma _x^{[s]}`$ ($`\sigma _z^{[s]}`$) have all eigenvalues $`\pm 1`$. In a classical error correcting code, correction proceeds by measuring the syndrome, which is done as follows. A parity check matrix $`H`$ of a code $`C`$ is a basis of the dual vector space $`C^{}`$. Suppose that we transmit a codeword $`v`$, which acquires errors to become $`w=v+ϵ`$. The $`k`$’th row $`r_k`$ of the matrix $`H`$ determines the $`k`$’th bit of the syndrome for $`w`$, namely $`r_kw\text{ (mod 2)}`$. The full syndrome is thus $`Hw`$. If the syndrome is $`0`$, then $`wC`$. Otherwise, the most likely value of the error $`ϵ`$ can be calculated from the syndrome . In our quantum CSS code, we need to correct both bit and phase errors. Let $`H_1`$ be a parity check matrix for the code $`C_1`$, and $`H_2`$ one for the code $`C_2^{}`$. To calculate the syndrome for bit flips, we measure the eigenvalue of $`\sigma _z^{[r]}`$ for each row $`rH_1`$ ($`1`$’s and $`1`$’s of the eigenvalue correspond to 1’s and 0’s of the syndrome). To calculate the syndrome for phase flips, we measure the eigenvalue of $`\sigma _x^{[r]}`$ for each row $`rH_2`$. This lets us correct both bit and phase flips, and if we can correct up to $`t`$ of each of these types of errors, we can also correct arbitrary errors on up to $`t`$ qubits . The useful property of CSS codes for demonstrating the security of BB84 is that the error correction for the phases is decoupled from that for the bit values, as shown above. General quantum stabilizer codes can similarly be turned into key distribution protocols, but these appear to require a quantum computer to implement. If one requires that a CSS code correct all errors on at most $`t=\delta n`$ qubits, the best codes that we know exist satisfy the quantum Gilbert-Varshamov bound. As the block length $`n`$ goes to infinity, these codes asymptotically protect against $`\delta n`$ bit errors and $`\delta n`$ phase errors, and encode $`[12H(2\delta )]n`$ qubits, where $`H`$ is the binary Shannon entropy $`H(p)=p\mathrm{log}_2(p)(1p)\mathrm{log}_2(1p)`$. In practice, it is better to only require that random errors are corrected with high probability. In this case, codes exist that correct $`\delta n`$ random phase errors and $`\delta n`$ random bit errors, and which encode $`[12H(\delta )]n`$ qubits. We also need a description of the Bell basis. These are the four maximally entangled states $`\mathrm{\Psi }^\pm ={\displaystyle \frac{1}{\sqrt{2}}}(|\mathrm{\hspace{0.17em}01}\pm |\mathrm{\hspace{0.17em}10}),\mathrm{\Phi }^\pm ={\displaystyle \frac{1}{\sqrt{2}}}(|\mathrm{\hspace{0.17em}00}\pm |\mathrm{\hspace{0.17em}11}),`$ which form an orthogonal basis for the quantum state space of two qubits. Finally, we introduce a class of quantum error correcting codes equivalent to $`Q`$, and parameterized by two $`n`$-bit binary vectors $`x`$ and $`z`$. Suppose that $`Q`$ is determined as above by $`C_1`$ and $`C_2`$. Then $`Q_{x,z}`$ has basis vectors indexed by cosets of $`C_2`$ in $`C_1`$, and for $`vC_1`$, the corresponding codeword is $$v\frac{1}{|C_2|^{1/2}}\underset{wC_2}{}(1)^{zw}|x+v+w.$$ (2) Quantum error correcting codes and entanglement purification protocols are closely connected ; we now describe the entanglement purification protocol corresponding to the CSS code $`Q`$. For now, we assume that the codes $`C_1`$ and $`C_2^{}`$ correct up to $`t`$ errors and that $`Q`$ encodes $`m`$ qubits in $`n`$ qubits. Suppose Alice and Bob share $`n`$ pairs of qubits in a state close to $`(\mathrm{\Phi }^+)^n`$. For the entanglement purification protocol, Alice and Bob separately measure the eigenvalues of $`\sigma _z^{[r]}`$ for each row $`rH_1`$ and $`\sigma _x^{[r^{}]}`$ for each row $`r^{}H_2`$. Note that for these measurements to be performable simultaneously, they must all commute; $`\sigma _z^{[r]}`$ and $`\sigma _x^{[r^{}]}`$ commute because the vector spaces $`C_1^{}`$ and $`C_2`$ are orthogonal. If Alice and Bob start with $`n`$ perfect EPR pairs, measuring $`\sigma _z^{[r]}`$ for $`rH_1`$ and $`\sigma _x^{[r^{}]}`$ for $`r^{}H_2`$ projects each of their states onto the code subspace $`Q_{x,z}`$, where $`x`$ and $`z`$ are any binary vectors with $`H_1x`$ and $`H_2z`$ equal to the measured bit and phase syndromes, respectively. After projection, the state is $`(\mathrm{\Phi }^+)^m`$ encoded by $`Q_{x,z}`$. Now, suppose that Alice and Bob start with a state close to $`(\mathrm{\Phi }^+)^n`$. To be specific, suppose that all their EPR pairs are in the Bell basis, with $`t`$ or fewer bit flips ($`\mathrm{\Psi }^+`$ or $`\mathrm{\Psi }^{}`$ pairs) and $`t`$ or fewer phase flips ($`\mathrm{\Phi }^{}`$ or $`\mathrm{\Psi }^{}`$ pairs). If Alice and Bob compare their measurements of $`\sigma _z^{[r]}`$ ($`\sigma _x^{[r]}`$), the rows $`r`$ for which these measurements disagree give the bits which are $`1`$ in the bit (phase) syndromes. From these syndromes, Alice and Bob can compute the locations of the bit and the phase flips, can correct these errors, and can then decode $`Q_{x,z}`$ to obtain $`m`$ perfect EPR pairs. We will show that the following is a secure quantum key distribution protocol. Protocol 1: Modified Lo-Chau * Alice creates $`2n`$ EPR pairs in the state $`(\mathrm{\Phi }^+)^n`$. * Alice selects a random $`2n`$ bit string $`b`$, and performs a Hadamard transform on the second half of each EPR pair for which $`b`$ is 1. * Alice sends the second half of each EPR pair to Bob. * Bob receives the qubits and publicly announces this fact. * Alice selects $`n`$ of the $`2n`$ encoded EPR pairs to serve as check bits to test for Eve’s interference. * Alice announces the bit string $`b`$, and which $`n`$ EPR pairs are to be check bits. * Bob performs Hadamards on the qubits where $`b`$ is $`1`$. * Alice and Bob each measure their halves of the $`n`$ check EPR pairs in the $`|\mathrm{\hspace{0.17em}0}`$, $`|\mathrm{\hspace{0.17em}1}`$ basis and share the results. If too many of these measurements disagree, they abort the protocol. * Alice and Bob make the measurements on their code qubits of $`\sigma _z^{[r]}`$ for each row $`rH_1`$ and $`\sigma _x^{[r]}`$ for each row $`rH_2`$. Alice and Bob share the results, compute the syndromes for bit and phase flips, and then transform their state so as to obtain $`m`$ nearly perfect EPR pairs. * Alice and Bob measure the EPR pairs in the $`|\mathrm{\hspace{0.17em}0}`$, $`|\mathrm{\hspace{0.17em}1}`$ basis to obtain a shared secret key. We now show that this protocol works. Namely, we show that the probability is exponentially small that Alice and Bob agree on a key about which Eve can obtain more than an exponentially small amount of information. We need a result of Lo and Chau that if Alice and Bob share a state having fidelity $`12^s`$ with $`(\mathrm{\Phi }^+)^m`$, then Eve’s mutual information with the key is at most $`2^c+2^{O(2s)}`$ where $`c=s\mathrm{log}_2(2m+s+1/\mathrm{log}_e2)`$. For the proof, we use an argument based on one from Lo and Chau . Let us calculate the probability that the test on the check bits succeeds while the entanglement purification on the code bits fails. We do this by considering the measurement that projects each of the EPR pairs onto the Bell basis. We first consider the check bits. Note that for the EPR pairs where $`b=1`$, Alice and Bob are effectively measuring them in the $`|+`$, $`|`$ basis rather than the $`|\mathrm{\hspace{0.17em}0}`$, $`|\mathrm{\hspace{0.17em}1}`$ basis. Now, observe that $`|\mathrm{\Psi }^+\mathrm{\Psi }^+|+|\mathrm{\Psi }^{}\mathrm{\Psi }^{}|`$ $`=`$ $`|\mathrm{\hspace{0.17em}01}\mathrm{\hspace{0.17em}01}|+|\mathrm{\hspace{0.17em}10}\mathrm{\hspace{0.17em}10}|,`$ $`|\mathrm{\Phi }^{}\mathrm{\Phi }^{}|+|\mathrm{\Psi }^{}\mathrm{\Psi }^{}|`$ $`=`$ $`|++|+|++|.`$ These relations show that the rates of bit flip errors and of phase flip errors that Alice and Bob estimate from their measurements on check bits are the same as they would have estimated using the Bell basis measurement. We next consider the measurements on the code bits. We want to show that the purification protocol applied to $`n`$ pairs produces a state that is close to the encoded $`(\mathrm{\Phi }^+)^m`$. The purification protocol succeeds perfectly acting on the space spanned by Bell pairs that differ from $`(\mathrm{\Phi }^+)^n`$ by $`t`$ or fewer bit flip errors and by $`t`$ or fewer phase flips errors. Let $`\mathrm{\Pi }`$ denote the projection onto this space. Then if the protocol is applied to an initial density operator $`\rho `$ of the $`n`$ pairs, it can be shown that the final density operator $`\rho ^{}`$ approximates $`(\mathrm{\Phi }^+)^m`$ with fidelity $$F(\mathrm{\Phi }^+)^m|\rho ^{}|(\mathrm{\Phi }^+)^m\mathrm{tr}\left(\mathrm{\Pi }\rho \right).$$ (3) Hence the fidelity is at least as large as the probability that $`t`$ or fewer bit flip errors and $`t`$ or fewer phase flip errors would have been found, if the Bell measurement had been performed on all $`n`$ pairs. Now, when Eve has access to the qubits, she does not yet know which qubits are check qubits and which are code qubits, so she cannot treat them differently. The check qubits that Alice and Bob measure thus behave like a classical random sample of the qubits. We are then able to use the measured error rates in a classical probability estimate; we find that probability of obtaining more than $`\delta n`$ bit (phase) errors on the code bits and fewer than $`(\delta ϵ)n`$ errors on the check bits is asymptotically less than $`\mathrm{exp}[\frac{1}{4}ϵ^2n/(\delta \delta ^2)]`$. We conclude that if Alice and Bob have greater than an exponentially small probability of passing the test, then the fidelity of Alice and Bob’s state with $`(\mathrm{\Phi }^+)^m`$ is exponentially close to 1. We now show how to turn this Lo-Chau type protocol into a quantum error-correcting code protocol. Observe first that it does not matter whether Alice measures her check bits before or after she transmits half of each EPR pair to Bob, and similarly that it does not matter whether she measures the syndrome before or after this transmission. If she measures the check bits first, this is the same as choosing a random one of $`|\mathrm{\hspace{0.17em}0}`$, $`|\mathrm{\hspace{0.17em}1}`$. If she also measures the syndrome first, this is equivalent to transmitting $`m`$ halves of EPR pairs encoded by the CSS code $`Q_{x,z}`$ for two random vectors $`x`$, $`z𝐅_2^n`$. The vector $`x`$ is determined by the syndrome measurements $`\sigma _z^{[r]}`$ for rows $`rH_1`$, and similarly for $`z`$. Alice can also measure her half of the encoded EPR pairs before or after transmission. If she measures them first, this is the same as choosing a random key $`k`$ and encoding $`k`$ using $`Q_{x,z}`$. We thus obtain the following equivalent protocol. Protocol 2: CSS Codes * Alice creates $`n`$ random check bits, a random $`m`$-bit key $`k`$, and a random $`2n`$-bit string $`b`$. * Alice chooses $`n`$-bit strings $`x`$ and $`z`$ at random. * Alice encodes her key $`|k`$ using the CSS code $`Q_{x,z}`$ * Alice chooses $`n`$ positions (out of $`2n`$) and puts the check bits in these positions and the code bits in the remaining positions. * Alice applies a Hadamard transform to those qubits in the positions having 1 in $`b`$. * Alice sends the resulting state to Bob. Bob acknowledges receipt of the qubits. * Alice announces $`b`$, the positions of the check bits, the values of the check bits, and the $`x`$ and $`z`$ determining the code $`Q_{x,z}`$. * Bob performs Hadamards on the qubits where $`b`$ is $`1`$. * Bob checks whether too many of the check bits have been corrupted, and aborts the protocol if so. * Bob decodes the key bits and uses them for the key. Intuitively, the security of the protocol depends on the fact that for a sufficiently low error rate, a CSS code transmits the information encoded by it with very high fidelity, so that by the no-cloning principle very little information can leak to Eve. We now give the final argument that turns the above protocol into BB84. First note that, since all Bob cares about are the bit values of the encoded key, and the string $`z`$ is only used to correct the phase of the encoded qubits, Bob does not need $`z`$. This is why we use CSS codes: they decouple the phase correction from the bit correction. Let $`k^{}C_1`$ be a binary vector that is mapped by Eq. (2) to the encoded key. Since Bob never uses $`z`$, we can assume that Alice does not send it. Averaging over $`z`$, we see that Alice effectively sends the mixed state $`{\displaystyle \frac{1}{2^n|C_2|}}{\displaystyle \underset{z}{}}[{\displaystyle \underset{w_1,w_2C_2}{}}(1)^{(w_1+w_2)z}`$ (4) $`\times |k^{}+w_1+xk^{}+w_2+x|]`$ (5) $`={\displaystyle \frac{1}{|C_2|}}{\displaystyle \underset{wC_2}{}}|k^{}+w+xk^{}+w+x|,`$ (6) which is equivalently the mixture of states $`|k^{}+x+w`$ with $`w`$ chosen randomly in $`C_2`$. Let us now look at the protocol as a whole. The error correction information Alice gives Bob is $`x`$, and Alice sends $`|k^{}+x+w`$ over the quantum channel. Over many iterations of the algorithm, these are random variables chosen uniformly in $`𝐅_2^n`$ with the constraint that their difference $`k^{}+w`$ is in $`C_1`$. After Bob receives $`k^{}+w+x+ϵ`$, he subtracts $`x`$, and corrects the result to a codeword in $`C_1`$, which is almost certain to be $`k^{}+w`$. The key is the coset of $`k^{}+w`$ over $`C_2`$. In the BB84 protocol given below, Alice sends $`|v`$ to Bob, with error correction information $`u+v`$. These are again two random variables uniform in $`𝐅_2^n`$, with the constraint that $`uC_1`$. Bob obtains $`v+ϵ`$, subtracts $`u+v`$, and corrects the result to a codeword in $`C_1`$, which with high probability is $`u`$. The key is then the coset $`u+C_2`$. Thus, the two protocols are completely equivalent. Protocol 3: BB84 * Alice creates $`(4+\delta )n`$ random bits. * Alice chooses a random $`(4+\delta )n`$-bit string $`b`$. For each bit, she creates a state in the $`|\mathrm{\hspace{0.17em}0}`$, $`|\mathrm{\hspace{0.17em}1}`$ basis (if the corresponding bit of $`b`$ is $`0`$) or the $`|+`$, $`|`$ basis (if the bit of $`b`$ is $`1`$). * Alice sends the resulting qubits to Bob. * Bob receives the $`(4+\delta )n`$ qubits, measuring each in the $`|\mathrm{\hspace{0.17em}0}`$,$`|\mathrm{\hspace{0.17em}1}`$ or the $`|+`$,$`|`$ basis at random. * Alice announces $`b`$. * Bob discards any results where he measured a different basis than Alice prepared. With high probability, there are at least $`2n`$ bits left (if not, abort the protocol). Alice decides randomly on a set of $`2n`$ bits to use for the protocol, and chooses at random $`n`$ of these to be check bits. * Alice and Bob announce the values of their check bits. If too few of these values agree, they abort the protocol. * Alice announces $`u+v`$, where $`v`$ is the string consisting of the remaining non-check bits, and $`u`$ is a random codeword in $`C_1`$. * Bob subtracts $`u+v`$ from his code qubits, $`v+ϵ`$, and corrects the result, $`u+ϵ`$, to a codeword in $`C_1`$. * Alice and Bob use the coset of $`u+C_2`$ as the key. There are a few loose ends that need to be tied up. The protocol given above uses binary codes $`C_1`$ and $`C_2^{}`$ with large minimum distance, and thus can obtain rates given by the quantum Gilbert-Varshamov bound for CSS codes . To reach the better Shannon bound for CSS codes, we need to use codes for which a random small set of phase errors and bit errors can almost always be corrected. To prove that the protocol works in this case, we need to ensure that the errors are indeed random. We do this by adding a step where Alice scrambles the qubits using a random permutation $`\pi `$ before sending them to Bob, and a step after Bob acknowledges receiving the qubits where Alice sends $`\pi `$ to Bob and he unscrambles the qubits. This can work as long as the measured bit and phase error rates are less than 11%, the point at which the Shannon rate $`12H(\delta )`$ hits $`0`$. For a practical key distribution protocol we need the classical code $`C_1`$ to be efficiently decodeable. As is shown in , we can let $`C_2`$ be a random subcode of an efficiently decodeable code $`C_1`$, and with high probability obtain a good code $`C_2^{}`$. While known efficiently decodeable codes do not meet the Shannon bound, they come fairly close. A weakness in both the proof given in this paper and the proofs in is that they do not apply to imperfect sources; the sources must be perfect single-photon sources. A proof avoiding this difficulty was recently discovered by Michael Ben-Or ; it shows that any source sufficiently close to a single-photon source is still secure. However, most experimental quantum key distribution systems use weak coherent sources, and no currently known proof covers this case. The authors thank Michael Ben-Or, Eli Biham, Hoi-Kwong Lo, Dominic Mayers and Tal Mor for explanations of and informative discussions about their security proofs. We also thank Ike Chuang, Dan Gottesman, Alexei Kitaev and Mike Nielsen for their discussions and suggestions, which greatly improved this paper. Part of this research was done while PWS was visiting Caltech. This work has been supported in part by the Department of Energy under Grant No. DE-FG03-92-ER40701, and by DARPA through Caltech’s Quantum Information and Computation (QUIC) project administered by the Army Research Office.
warning/0003/astro-ph0003024.html
ar5iv
text
# Magnetic field strength from peaked synchrotron spectra ## 1 Introduction The evaluation of the magnetic field strength of radio sources may in principle be based on the fact that their synchrotron spectra exhibit a maximum, which arises due to self absorption in the emitting plasma (Rybicki & Lightman 1979). This necessitates the full knowledge of the electron<sup>1</sup><sup>1</sup>1Although not stated explicitly, all results of this paper are valid for electron and electron-positron plasmas alike. distribution. However, whereas the distribution in the optically thin regime can be inferred from the spectrum and is known to be of power law form, it remains uncertain in the optically thick regime, so that for the latter some assumption is needed. Obviously, the simplest alternative would be to assume an unbroken power law (cf. Pacholczyk 1970). In this article, the differences between an unbroken power law and some other physically more realistic distributions are investigated. Throughout, apart from the assumption of homogeneity and isotropy, the discussion will be kept as independent as possible from specific source parameters. In Section 2 the kinetic equations governing emission and absorption in a synchrotron plasma are reviewed, and two simple steady state solutions are given. They motivate the use of either a power law with index -3 or a thermal electron distribution in the optically thick regime. Section 3 deals with the magnetic field strengths one obtains under the assumption that the transition from the optically thick to the optically thin regime corresponds to the turnover energy in the electron distribution. Piecewise power laws and thermal distributions for low energies are discussed. In Section 4 the determination of the magnetic field strength from the maximum intensity of the spectrum is investigated. The uncertainty in the limiting brightness temperature is discussed briefly in Section 5. ## 2 Kinetic equations Treating synchrotron radiation as a special case of bremsstrahlung, we can see that synchrotron emission as a spontaneous process must be accompanied by the corresponding induced processes. Accordingly, it may be described by means of the Einstein coefficients, so that the Einstein relations $`A_{21}=\frac{2h\nu ^3}{c^2}B_{21}`$ and $`B_{12}=B_{21}`$ may be applied. In order to simplify, in the following we assume the electron energies $`ϵ`$ are highly relativistic with $`ϵ\mathrm{}\omega `$ ($`\omega `$ being the angular frequency of an emitted photon) and that both the magnetic field and the electron distribution may be treated as isotropic. Replacing $`A_{21}`$ by the power spectrum $`P(ϵ,\omega )`$ and employing the Einstein relations then yields the kinetic equations describing a synchrotron emitting plasma (McCray 1969, Norman & ter Haar 1974): $$\frac{W(\omega )}{t}=dϵ\overline{P}(ϵ,\omega )\left\{\frac{}{}f(ϵ)+\frac{\pi ^2c^3}{\omega ^2}ϵ^2\frac{}{ϵ}\left(\frac{f(ϵ)}{ϵ^2}\right)W(\omega )\right\}$$ $$\frac{f(ϵ)}{t}=\frac{}{ϵ}\left(d\omega \overline{P}(ϵ,\omega )\left\{f(ϵ)+\frac{\pi ^2c^3}{\omega ^2}ϵ^2\frac{}{ϵ}\left(\frac{f(ϵ)}{ϵ^2}\right)W(\omega )\right\}\right)$$ In these formulae, $`f`$ stands for the electron density per energy $`ϵ`$, $`W`$ for the photon energy density per angular frequency $`\omega `$. $`\overline{P}`$ is the angular average of $`P`$, i.e. $`\overline{P}=_{4\pi }Pd\mathrm{\Omega }`$. The kinetic equations obtained in this manner remain valid in case of a homogeneous magnetic field and an isotropic photon distribution. In the stationary case comparison of the kinetic equations and the transfer equation $`\frac{\mathrm{d}I_\nu }{\mathrm{d}s}=j_\nu \alpha _\nu I_\nu `$ with emission coefficient $`j_\nu `$, absorption coefficient $`\alpha _\nu `$, and path length $`s`$ yields an expression for $`\alpha _\nu `$: $$\alpha _\nu =dϵ\overline{P}(ϵ,\omega )\frac{\pi ^2c^2}{\omega ^2}ϵ^2\frac{}{ϵ}\left(\frac{f(ϵ)}{ϵ^2}\right)$$ (1) Turning now from these rather general considerations to the specific example of synchrotron radiation, one starts by defining $`x\omega /\omega _\mathrm{c}`$ and $`F(x)x_x^{\mathrm{}}K_{5/3}(\xi )d\xi `$, where $`\omega _\mathrm{c}3ϵ^2eB\mathrm{sin}\alpha /2m_\mathrm{e}^3c^5`$ with the elementary charge $`e`$, the electron mass $`m_\mathrm{e}`$ and the pitch angle $`\alpha `$ between magnetic field and electron velocity. $`K_{5/3}`$ constitutes the modified Bessel function of order 5/3. Using these definitions, one may express the synchrotron power spectrum as $`P(ϵ,\omega )=(\sqrt{3}e^3B\mathrm{sin}\alpha /2\pi m_\mathrm{e}c^2)^{1/2}F(x)`$ (Ginzburg & Syrovatskii 1965). Its angular average is given by $`\overline{P}(ϵ,\omega )=Q(\stackrel{~}{x})P(ϵ,\omega )|_{\mathrm{sin}\alpha =1}`$ with $`\stackrel{~}{x}x|_{\mathrm{sin}\alpha =1}`$, where for $$Q(\stackrel{~}{x})=\frac{1}{2F(\stackrel{~}{x})}_1^{+1}d\mu \sqrt{1\mu ^2}F\left(\frac{\stackrel{~}{x}}{\sqrt{1\mu ^2}}\right)$$ the approximation $`Q(\stackrel{~}{x})_{n=0}^3Q_n\stackrel{~}{x}^n0.8225330.181121\stackrel{~}{x}+0.0370118\stackrel{~}{x}^20.00316623\stackrel{~}{x}^3`$ can be used. (Because of $`\overline{P}(ϵ,\omega )P(ϵ,\omega )|_{\mathrm{sin}\alpha =1}`$, $`Q(\stackrel{~}{x})1`$, so that for $`\stackrel{~}{x}\mathrm{}`$ the approximation becomes arbitrarily inaccurate. However, $`Q(\stackrel{~}{x})`$ will always be part of an integrand containing $`\mathrm{e}^{\stackrel{~}{x}}`$, which renders this inaccuracy irrelevant.) Concerning stationary distributions, two important solutions of the kinetic equations can be stated (Norman & ter Haar 1974). Firstly, the distributions may be those of thermal equilibrium in the Rayleigh-Jeans limit, i.e. $`f(ϵ)=f_0\mathrm{exp}\left(ϵ/\pi ^2c^3\mathrm{\Omega }_0\right)`$, $`W(\omega )=\mathrm{\Omega }_0\omega ^2`$, where $`\varphi _0`$ and $`\mathrm{\Omega }_0`$ are arbitrary positive constants. Secondly, the only viable power law for the electron distribution has the index -3, so that $`f(ϵ)=f_0ϵ^3`$. The corresponding photon distribution has the form $`W(\omega )=\frac{1}{5\pi ^2}\sqrt{\frac{2m_\mathrm{e}^3}{3ceB}}\omega ^{5/2}`$. Finally, one may note that the form of the kinetic equations precludes the use of the delta approximation $`\overline{P}(ϵ,\omega )=C_1ϵ^2\delta (\omega C_2ϵ^2)`$ ($`C_1`$, $`C_2`$ being constants) for the power spectrum, as, despite of yielding the correct photon distributions, this approximation allows any electron distribution, which contradicts the fact that only one power law index is allowed. ## 3 Magnetic field strength A typical synchrotron spectrum contains a turnover, marking the transition from optically thick to optically thin emission. Whereas, from observation, the optically thin emission (and thus the electron spectrum for sufficiently high energies) can generally be taken to be a power law with some index $`p_2`$, the precise form of the optically thick spectrum is hard to detect. Therefore the electron spectrum is taken to be of the form $$f(ϵ)=\{\begin{array}{ccc}F(ϵ)\hfill & ,& ϵ<ϵ_\mathrm{T}\hfill \\ f_0ϵ^{p_2}\hfill & ,& ϵϵ_\mathrm{T}\hfill \end{array}$$ where various functions $`F`$ will be discussed. The simplest choice is to assume an unbroken power law spectrum, i.e. $`F(ϵ)=f_0ϵ^{p_2}`$. Quantities relating to this choice are denoted by the index a. Inserting $`f_\mathrm{a}(ϵ)`$ and the (averaged) synchrotron power spectrum $`\overline{P}(ϵ,\omega )`$ into equation (1), one obtains, using the relation $`x^\mu F(x)dx=\frac{2^{\mu +1}}{\mu +2}\mathrm{\Gamma }\left(\frac{\mu }{2}+\frac{7}{3}\right)\mathrm{\Gamma }\left(\frac{\mu }{2}+\frac{2}{3}\right)`$ with $`\mu >\frac{4}{3}`$ (Westfold 1959), $$\alpha _\nu =\underset{n=0}{\overset{3}{}}\alpha _0f_0R_\mathrm{a}^{\frac{p_2}{2}}\nu ^{\frac{p_2}{2}2}h(p_2,n),$$ (2) where $`\alpha _0\frac{\sqrt{3}e^3B}{8\pi m_\mathrm{e}}`$, $`R\frac{3eB}{4\pi m_\mathrm{e}^3c^5}`$ and $$h(p,n)Q_n\left(1\frac{p}{2}\right)\frac{2^{n\frac{p}{2}}}{n+1\frac{p}{2}}\mathrm{\Gamma }\left(\xi _1\right)\mathrm{\Gamma }\left(\xi _2\right),$$ where the $`\xi _i`$ are defined as $`\xi _1(6n+223p)/12`$ and $`\xi _2(6n+23p)/12`$. Note that equation (2) can easily be solved for the magnetic field strength $`B`$. A second, slightly more general choice for $`f`$, which will be indicated by the index b, is a piecewise power law, so that $`F(ϵ)=f_0ϵ_\mathrm{T}^{p_2p_1}ϵ^{p_1}`$. Here the constant $`ϵ_\mathrm{T}^{p_2p_1}`$ is implied by demanding continuity in $`ϵ_\mathrm{T}`$. Using the approximation $`F(x)F_0\stackrel{~}{x}^\varphi \mathrm{e}^{\stackrel{~}{x}}=1.8\stackrel{~}{x}^{0.3}\mathrm{e}^{\stackrel{~}{x}}`$ (Melrose 1980) together with $`\stackrel{~}{x}(ϵ,\nu )x|_{\mathrm{sin}\alpha =1}=\nu /Rϵ^2`$ and $$g(p,n,\stackrel{~}{x}_1,\stackrel{~}{x}_2)F_0Q_n\left\{\zeta _3\mathrm{\Gamma }_{\stackrel{~}{x}_1}^{\stackrel{~}{x}_2}\left(\zeta _1\right)\mathrm{\Gamma }_{\stackrel{~}{x}_1}^{\stackrel{~}{x}_2}\left(\zeta _2\right)\right\}$$ (where $`\zeta _1n+\varphi p/2`$, $`\zeta _2n+\varphi p/2+1`$ and $`\zeta _3n+\varphi 1`$) one gets for the ratio $`\alpha _{\nu _1,\mathrm{b}}/\alpha _{\nu _2,\mathrm{a}}`$ of the absorption coefficients computed with $`f_\mathrm{a}`$ and $`f_\mathrm{b}`$ $$\frac{\alpha _{\nu _1,\mathrm{b}}}{\alpha _{\nu _2,\mathrm{a}}}=\frac{_{n=0}^3\left\{\alpha _{\mathrm{den}}(\nu _1,ϵ_{\mathrm{T},\mathrm{b}},p_1,p_2,n)\right\}}{_{n=0}^3\alpha _{0,\mathrm{a}}f_0R_\mathrm{a}^{\frac{p_2}{2}}\nu _2^{\frac{p_2}{2}2}h(p_2,n)}$$ (3) where $`\alpha _{\mathrm{den}}(\nu _1,ϵ_{\mathrm{T},\mathrm{b}},p_1,p_2,n)`$ $`=\alpha _{0,\mathrm{b}}f_0R_\mathrm{b}^{\frac{p_2}{2}}\nu _1^{\frac{p_2}{2}2}g(p_2,n,0,\stackrel{~}{x}(\nu _1,ϵ_{\mathrm{T},\mathrm{b}}))`$ $`+\alpha _{0,\mathrm{b}}f_0ϵ^{p_2p_1}R_\mathrm{b}^{\frac{p_1}{2}}\nu _1^{\frac{p_1}{2}2}g(p_1,n,\stackrel{~}{x}(\nu _1,ϵ_{\mathrm{T},\mathrm{b}}),\mathrm{}).`$ Now the synchrotron power spectrum takes its maximum at the frequency $`\nu _{\mathrm{mp}}(ϵ)=0.29\frac{\pi }{4}Rϵ^2`$, so that $`\stackrel{~}{x}(\nu _{\mathrm{mp}}(ϵ_\mathrm{T}),ϵ_\mathrm{T})=0.29\frac{\pi }{4}Rϵ_\mathrm{T}^2/Rϵ_\mathrm{T}^2=0.29\frac{\pi }{4}`$. As $`ϵ_\mathrm{T}`$ constitutes the transition between the parts of the electron spectrum yielding optically thin and optically thick radiation, it is reasonable to assume that the optical depth corresponding to $`\nu _{\mathrm{mp}}(ϵ_\mathrm{T})`$ is approximately 1, i.e. that for a homogeneous plasma $`\alpha _{\nu _{\mathrm{mp}}(ϵ_\mathrm{T})}L1`$, where $`L`$ denotes the spatial extension of the plasma. (In the case of an unbroken power law, this condition is used to define $`ϵ_\mathrm{T}`$.) Hence $`1\alpha _{\nu _{\mathrm{mp}}(ϵ_\mathrm{T}),\mathrm{b}}/\alpha _{\nu _{\mathrm{mp}}(ϵ_\mathrm{T}),\mathrm{a}}`$. If one allows for different turnover energies, this may be generalized to $`1\alpha _{\nu _{\mathrm{mp}}(ϵ_{\mathrm{T},\mathrm{b}}),\mathrm{b}}/\alpha _{\nu _{\mathrm{mp}}(ϵ_{\mathrm{T},\mathrm{a}}),\mathrm{a}}`$, and using equation (3) and $`\stackrel{~}{x}(\nu _{\mathrm{mp}}(ϵ_\mathrm{T}),ϵ_\mathrm{T})=0.29\frac{\pi }{4}`$ together with the ratio $`r_\mathrm{b}ϵ_{\mathrm{T},\mathrm{b}}/ϵ_{\mathrm{T},\mathrm{a}}`$ one can obtain the ratio $`b_\mathrm{b}=B_\mathrm{b}/B_\mathrm{a}`$ of the magnetic field strengths for $`f_\mathrm{a}`$ and $`f_\mathrm{b}`$, respectively: $$b_\mathrm{b}=r_\mathrm{b}^{p_24}\frac{_{n=0}^3\left\{\gamma _1+\left(\frac{0.29\pi }{4}\right)^{\frac{1}{2}(p_1p_2)}\gamma _2\right\}}{_{n=0}^3h(p_2,n)}$$ Here the $`\gamma _i`$ denote $`\gamma _1=g(p_2,n,0,\frac{0.29}{4}\pi )`$ and $`\gamma _2=g(p_1,n,\frac{0.29}{4}\pi ,\mathrm{})`$. If the turnover frequency is independent of the electron spectrum (i.e. if $`\nu _{\mathrm{mp},\mathrm{a}}(ϵ_{\mathrm{T},\mathrm{a}})=\nu _{\mathrm{mp},\mathrm{b}}(ϵ_{\mathrm{T},\mathrm{b}})`$), the relation $`br_\mathrm{b}^2=1`$ holds valid. Hence we assume that $`br_\mathrm{b}^2=N`$ with some arbitrary, but constant $`N`$. In view of the discussion at the end of the previous section, one should take $`p_1=3`$. For $`N`$=0.25, 1, and 2, the plots of $`b_\mathrm{b}`$ versus $`p_2`$ corresponding to this choice of $`p_1`$ are shown in Fig. 1. Note that $`b_\mathrm{b}(p_2=3)1`$ for $`N=1`$ results from the use of the approximation $`F(x)1.8x^{0.3}e^x`$ in computing $`\alpha _{\nu ,\mathrm{b}}`$. A simple generalization of $`f_\mathrm{b}`$ is achieved by dropping the assumption that the electron spectrum extends down to zero energy (which, due to the finite electron rest energy, is questionable anyway), i.e. by taking a spectrum $`f_\mathrm{b}^{}(ϵ)f_\mathrm{b}(ϵ)\mathrm{\Theta }(ϵϵ_0)`$ rather than $`f_\mathrm{b}(ϵ)`$, where $`ϵ_0>0`$ and $`\mathrm{\Theta }`$ denotes the Heaviside function. This merely involves replacing $`\stackrel{~}{x}_2=\mathrm{}`$ in the argument of $`g(p,n,\stackrel{~}{x}_1,\stackrel{~}{x}_2=\mathrm{})`$ by $`\stackrel{~}{x}(\nu ,ϵ_0)`$ with the cutoff energy $`ϵ_0`$. If one introduces the ratio $`eϵ_{\mathrm{T},\mathrm{b}}/ϵ_0`$, one obtains $`\stackrel{~}{x}(\nu _{\mathrm{mp}}(ϵ_{\mathrm{T},\mathrm{b}}),ϵ_0)=0.29\frac{\pi }{4}e^2`$ and thus for the ratio $`b_\mathrm{b}^{}=B_\mathrm{b}^{}/B_\mathrm{a}`$ $$b_\mathrm{b}^{}=r_\mathrm{b}^{p_24}\frac{_{n=0}^3\left\{\gamma _1+\left(\frac{0.29\pi }{4}\right)^{\frac{1}{2}(p_1p_2)}\overline{\gamma }_2\right\}}{_{n=0}^3h(p_2,n)},$$ where $`\overline{\gamma }_2g(p_1,n,\frac{0.29}{4}\pi ,0.29\frac{\pi }{4}e^2)`$ and $`\gamma _1`$ is defined as above. Again, $`p_1`$ should be set to -3. The plots of $`b_\mathrm{b}^{}`$ as a function of $`e`$ resulting from this formula are given in Fig. 2 for several values of $`p_2`$ and $`N=1`$. As expected, in the limit of $`e\mathrm{}`$ (no energy cutoff) $`b_\mathrm{b}^{}`$ approaches $`b_\mathrm{b}`$; for values of $`e`$ exceeding 10, the two are essentially indistinguishable. Referring to the examples in the previous section, it is tempting to assume, as a further possibility for the electron spectrum, a thermal distribution for low energies, i.e. to choose $`F(ϵ)=f_0\mathrm{e}^{\beta ϵ_\mathrm{T}}ϵ_\mathrm{T}^{p_22}ϵ^2\mathrm{e}^{\beta ϵ}`$, where $`\beta 1/k_\mathrm{B}T`$ with the Boltzmann constant $`k_\mathrm{B}`$ and the temperature $`T`$. One can see easily that continuity in $`ϵ_\mathrm{T}`$ is ensured. The quantities relating to this choice of $`f(ϵ)`$ are marked by the index c. Inserting $`f_\mathrm{c}(ϵ)`$ into equation (1) and using $`\stackrel{~}{x}(\nu _{\mathrm{mp}}(ϵ_{\mathrm{T},\mathrm{c}}),ϵ_{\mathrm{T},\mathrm{c}})=0.29\frac{\pi }{4}`$ yields $`\alpha _{\nu ,\mathrm{c}}(\overline{\nu })={\displaystyle \underset{n=0}{\overset{3}{}}}\{\alpha _{0,c}f_0R_\mathrm{c}^{\frac{p_2}{2}}\overline{\nu }^{\frac{p_2}{2}2}\gamma _1`$ $`+\alpha _{0,\mathrm{c}}f_0\mathrm{\Delta }(p_1,p_2)\beta \mathrm{e}^{\beta ϵ_{\mathrm{T},\mathrm{c}}}\stackrel{~}{g}(\beta ϵ_{\mathrm{T},\mathrm{c}},n)\},`$ where $`\mathrm{\Delta }(p_1,p_2)\left(\frac{0.29\pi }{4}\right)^{\frac{p_2}{2}+1}R_\mathrm{c}^{\frac{p_2}{2}\frac{1}{2}}\overline{\nu }^{\frac{p_1}{2}\frac{3}{2}}`$, $`\overline{\nu }\nu _{\mathrm{mp}}(ϵ_{\mathrm{T},\mathrm{c}})`$, and $$\stackrel{~}{g}(u,n)=\frac{1}{2}F_0Q_n_{\frac{0.29}{4}\pi }^{\mathrm{}}d\stackrel{~}{x}\stackrel{~}{x}^{n+\varphi \frac{5}{2}}\mathrm{e}^{\stackrel{~}{x}}\mathrm{e}^{\sqrt{\frac{0.29}{4}\pi }u\frac{1}{\sqrt{x}}}.$$ Now in order to simplify the discussion, we assume $`N_{\mathrm{th}}`$ to be minimized. This assumption is motivated by the low photon density in the optically thin part of jet spectra, as implied by polarisation measurements (Jones 1988). Hence one demands $`\frac{\mathrm{d}N_{\mathrm{th}}}{\mathrm{d}ϵ_{\mathrm{T},\mathrm{c}}}=0`$, which together with $`u\beta ϵ_{\mathrm{T},\mathrm{c}}`$ leads to $`0=2(p_22)(\mathrm{e}^u1)u^3+2(\mathrm{e}^u(p_21))u^2p_2u^1.`$ Let $`u_{\mathrm{min}}=u_{\mathrm{min}}(p_2)`$ be the solution of this implicit equation. Then using Eqs. 2 and 3 and defining $`\stackrel{~}{\gamma }\stackrel{~}{g}(u_{\mathrm{min}},n)`$, one obtains for the ratio $`b_\mathrm{c}B_\mathrm{c}/B_\mathrm{a}`$ $$b_\mathrm{c}=r_\mathrm{c}^{p_24}\frac{_{n=0}^3\left\{\gamma _1+\left(\frac{0.29\pi }{4}\right)^{\frac{p_2}{2}+\frac{3}{2}}u_{\mathrm{min}}\mathrm{e}^{u_{\mathrm{min}}}\stackrel{~}{\gamma }\right\}}{_{n=0}^3h(p_2,n)},$$ where $`r_\mathrm{c}=ϵ_{\mathrm{T},\mathrm{c}}/ϵ_{\mathrm{T},\mathrm{a}}`$, the analogue of $`r_\mathrm{b}`$, is assumed to be given by $`br_\mathrm{c}^2=N`$ with some constant $`N`$. The plots of $`b_\mathrm{c}`$ versus $`p_2`$ for several values of $`N`$ are shown in Fig. 3. Note that the restriction to values of $`p_2`$ less than -1 is necessitated by the demand that $`u_{\mathrm{min}}`$ be positive. ## 4 Maximum intensity So far, it has been assumed that the optical depth of the frequency corresponding to the turnover energy is equal to 1. However, from an observational point of view, it is clearly more desirable to deduce the magnetic field from the position of the maximum intensity. In order to achieve this aim, one assumes the radiative transfer to be one-dimensional and starts from the general formula for the intensity $`I_\nu `$ of a homogeneous source (Pacholczyk 1970), $$I_\nu =\frac{j_\nu }{\alpha _\nu }(1\mathrm{e}^{\alpha _\nu L}).$$ (4) Now for a piecewise power law with a lower cutoff energy $`e^1ϵ_\mathrm{T}`$, using the notations introduced in the previous section, one may write the absorption coefficient as $`\alpha _\nu =_{i=1}^2_{n=0}^3\alpha _0A_iR^{\frac{p_i}{2}}\nu ^{\frac{p_i}{2}2}g(p_i,n,\stackrel{~}{x}_{1i},\stackrel{~}{x}_{2i})`$, where $`\stackrel{~}{x}_{11}=\stackrel{~}{x}(\nu ,e^1ϵ_\mathrm{T})`$, $`\stackrel{~}{x}_{12}=\stackrel{~}{x}_{21}=\stackrel{~}{x}(\nu ,ϵ_\mathrm{T})`$, $`\stackrel{~}{x}_{22}=\stackrel{~}{x}(\nu ,\mathrm{})=0`$, $`A_1=f_0ϵ_\mathrm{T}^{p_2p_1}`$, and $`A_2=f_0`$. If one introduces the turnover frequency $`\nu _0\nu _{\mathrm{mp}}(ϵ_\mathrm{T})`$, the $`\stackrel{~}{x}_{ik}`$ become functions of this frequency only, i.e. they may be considered to be independent of the magnetic field. Accordingly, $`\alpha _\nu `$ takes the form $`\alpha _{\nu ,\nu _0}=_{i=1}^2f_0F_{\alpha i}(\nu ,\nu _0)G_{\alpha i}(B)`$ with polynomial functions $`G_{\alpha i}(B)=B^{\kappa _i}`$. However, from $`\alpha _0RB`$ and $`A_1ϵ_\mathrm{T}^{p_2p_1}\nu _0^{1/2}B^{(p_2p_1)/2}`$ it follows that $`\kappa _1=\kappa _2=1\frac{p_2}{2}\kappa `$, yielding $`\alpha _\nu =G_\alpha (\nu ,\nu _0)B^\kappa `$ and hence $`\frac{}{\nu }\alpha _\nu =f_0F_\alpha (\nu ,\nu _0)B^\kappa `$. Similarly one obtains $`j_\nu =f_0F_j(\nu ,\nu _0)B^\lambda `$ and $`\frac{}{\nu }j_\nu =f_0F_j(\nu ,\nu _0)B^\lambda `$ (with $`\lambda =\frac{1}{2}(1p_2`$)) for the emissivity and its derivative. The intensity takes its maximum when its derivative vanishes. Therefore, by differentiating equation (4) one obtains the condition $`0=C_1(\nu _{\mathrm{max}},\nu _0)(1\mathrm{e}^{C_2(\nu _{\mathrm{max}},\nu _0)x})`$ (5) $`+C_3((\nu _{\mathrm{max}},\nu _0))x\mathrm{e}^{C_2(\nu _{\mathrm{max}},\nu _0)x},`$ where $`xf_0LB^\kappa `$, $`C_1=F_j/F_\alpha (F_j/F_\alpha ^2)F_\alpha `$, $`C_2=F_\alpha `$, and $`C_3=(F_j/F_\alpha )F_\alpha `$. Note that this equation is formally independent of both the source extension and the overall electron density (i.e. of $`f_0`$). Only when computing the magnetic field strength by means of $`B=(x/L)^{1/\kappa }`$ these parameters reenter. Clearly, for two models with the same values of $`Lf_0`$, $`\nu _{\mathrm{max}}`$, and $`\nu _0`$, the ratio of the magnetic field strengths to the power of $`\kappa `$ equals that of the $`x`$ obtained from equation (5). Hence, it is straightforward to use the ratios $`b_\mathrm{b}`$ and $`b_\mathrm{b}^{}`$ as defined in the previous section. Incidentally, this leads to a further simplification: Due to the fact that $`\stackrel{~}{x}(\nu ,e^1ϵ_\mathrm{T})`$ is a function of $`y\nu /\nu _0`$ and that $`A_1ϵ_\mathrm{T}^{(p_2p_1)}\nu _0^{\frac{1}{2}(p_2p_1)}`$, the functions $`G`$ defined above may be written as $`G_\alpha (\nu ,\nu _0)=\nu _0^{\frac{p_2}{2}2}\overline{G}_\alpha (y)`$, $`G_\alpha (\nu ,\nu _0)=\nu _0^{\frac{p_2}{2}3}\overline{G}_\alpha (y)`$, $`G_j(\nu ,\nu _0)=\nu _0^{\frac{p_2}{2}+\frac{1}{2}}\overline{G}_j(y)`$, and $`G_j(\nu ,\nu _0)=\nu _0^{\frac{p_2}{2}\frac{1}{2}}\overline{G}_j(y)`$, which implies $`C_1(\nu ,\nu _0)=\nu _0^{\frac{3}{2}}\overline{C}_1(y)`$, $`C_2(\nu ,\nu _0)=\nu _0^{\frac{p_2}{2}2}\overline{C}_2(y)`$, and $`C_3(\nu ,\nu _0)=\nu _0^{\frac{p_2}{2}\frac{1}{2}}\overline{C}(y)`$. Defining $`\overline{x}(y)`$ as the solution of $$\overline{C}_1(y)(1\mathrm{e}^{\overline{C}_2(y)\overline{x}(y)})+\overline{C}_3(y)\overline{x}(y)\mathrm{e}^{\overline{C}_2(y)\overline{x}(y)}=0$$ and multiplying both sides of this equation by $`\nu _0^{3/2}`$, one can see easily that equation (5) is solved by $`x(\nu _{\mathrm{max}},\nu _0)=\nu _0^{\frac{p_2}{2}+2}\overline{x}(y_{\mathrm{max}})`$, where $`y_{\mathrm{max}}\nu _{\mathrm{max}}/\nu _0`$. Accordingly, the ratios $`x_\mathrm{b}/x_\mathrm{a}`$ and $`x_\mathrm{b}^{}/x_\mathrm{a}`$ of the $`x`$ computed for the electron distributions a, b and $`\mathrm{b}^{}`$ of the previous section depend on $`y_{\mathrm{max}}`$ only. However, (assuming $`L`$ and $`f_0`$ are the same for all distributions) these ratios clearly constitute the ratios of the respective magnetic field strengths to the power of $`\kappa `$, so that $`b_\mathrm{b}b_\mathrm{b}(y_{\mathrm{max}})=(x_\mathrm{b}/x_\mathrm{a})^{1/\kappa }`$, $`b_\mathrm{b}^{}b_\mathrm{b}^{}(y_{\mathrm{max}})=(x_\mathrm{b}^{}/x_\mathrm{a})^{1/\kappa }`$. Hence the magnetic field ratios can be computed by means of solving equation (5) for the distributions involved. For several values of $`p_2`$ and $`e`$ the results obtained numerically in this way are plotted in Figs. 4 to 6. Now although this choice is not motivated by the discussion of Section 2, it is instructive to examine an electron distribution with a positive power law index $`p_1`$ in the low energy regime. For the sake of definiteness, $`p_1=2`$ is taken as an example, which may be regarded as a crude approximation of the thermal distribution. One then obtains the ratios $`b_\mathrm{b}`$ plotted in Fig. 7. Clearly, the result resembles that for $`p_1=3`$ and $`e=2`$, which implies that the contribution of the low energy part of the electron distribution is rather small. $`x_\mathrm{a}`$ and thus $`B_\mathrm{a}`$ do not depend on $`\nu _0`$ (see equation (6) below). Hence, assuming some specific value for $`\nu _{\mathrm{max}}`$, the ratios $`b_i`$ are given by $`b_i(\nu _{\mathrm{max}}/\nu _0)=CB_i(\nu _{\mathrm{max}}/\nu _0)`$, where the constant $`C`$ is independent of the spectrum $`i`$. Therefore $`B_i(z_1)/B_j(z_2)=b_i(z_1)/b_j(z_2)`$ (where $`z\nu _{\mathrm{max}}/\nu _0)`$, which allows a direct comparison of Figs. 4 to 7. This clearly shows that the magnetic fields obtained for the various electron spectra discussed in these figures are of the same order of magnitude.<sup>2</sup><sup>2</sup>2Obviously, this is not true, if, in Figs. 5 to 7, $`\nu _{\mathrm{max}}/\nu _0`$ happens to be smaller than the value corresponding to the maximum of the relevant curve. In that case, however, due to the steepness of the curve, one should distrust the magnetic field determination anyway. Obviously, in order to get $`x_\mathrm{b}`$ (or $`x_\mathrm{b}^{}`$) instead of the ratio $`x_\mathrm{b}/x_\mathrm{a}`$ (or $`x_\mathrm{b}^{}/x_\mathrm{a}`$, respectively), one needs to know the value of $`x_\mathrm{a}`$. Now, as is evident from the above discussion, $`x_\mathrm{a}`$ must be of the form $`x_\mathrm{a}(\nu _{\mathrm{max}},\nu _0)=\nu _0^{\frac{p_2}{2}+2}\overline{x}(y_{\mathrm{max}})`$. However, concerning an unbroken power law, the turnover frequency has no physical meaning, as there is no turnover at all. Still, it may be retained as a formal parameter, so that the distribution is viewed as a piecewise power law, where the power happens to be the same for all pieces. This implies that for $`\nu _0`$ any value may be assumed. Choosing $`\nu _0=\nu _{\mathrm{max}}`$, one obtains $`y_{\mathrm{max}}=1`$ and thus $$x_\mathrm{a}=\overline{x}(1)\nu _{\mathrm{max}}^{\frac{p_2}{2}+2}.$$ (6) Finally, an important technical remark should be made: The aim of the foregoing discussion was to gain results depending on as few parameters as possible. In order to achieve this, the turnover frequency rather than energy was employed. However, in a given model, the latter rather than the former will be given. In this case the magnetic field ratio cannot simply be read off, and the field strength has to be inferred from demanding self-consistency: Assuming some specific values of $`L`$, $`f_0`$ and the magnetic field $`B`$, one can compute the turnover frequency from the corresponding energy and thus obtain the ratio $`x_\mathrm{b}/x_\mathrm{a}`$ (or $`x_\mathrm{b}^{}/x_\mathrm{a}`$). Together with the value of $`x_\mathrm{a}`$ taken from equation (6) one gets the value of $`x_\mathrm{b}`$ (or $`x_\mathrm{b}^{}`$), leading to the magnetic field strength, which obviously should equal the one originally assumed. ## 5 An example: peak brightness temperature As a straightforward example of using the results obtained so far, we discuss the maximum brightness temperature of a (synchrotron) source with an isotropic magnetic field (cf. Kellermann 1974). To do so, one starts by noting that the ratio of the synchrotron and (to first order) inverse Compton losses is given by $$\frac{P_\mathrm{c}}{P_{\mathrm{syn}}}=\frac{u_{\mathrm{rad}}}{u_B}=\frac{L/(\pi \rho ^2c)}{B^2/(8\pi )}=\frac{32\pi \xi _{\mathrm{geom}}R^2F_{\mathrm{max}}\nu _\mathrm{c}}{\rho ^2cB^2}.$$ Here, $`u_{\mathrm{rad}}`$ and $`u_B`$ denote the energy density of the radiation and the magnetic field, respectively, $`\rho `$ the radius of the source, $`R`$ its distance from the observer, and $`\nu _\mathrm{c}`$ the upper cut-off frequency. Clearly, $`\rho =R\theta /2`$ with the angular size $`\theta `$. The luminosity $`L`$ is approximated as $`L\xi _{\mathrm{geom}}4\pi R^2F_\mathrm{m}\nu _\mathrm{c}`$, where $`F_\mathrm{m}`$ is the observed maximum flux density and $`\xi _{\mathrm{geom}}`$ takes into account any deviation from isotropic radiation. The solid angle of the source (as seen from the observer) is given by $`\pi \theta ^2/4`$. Hence, assuming that the one-dimensional transfer equation (equation (4)) may be applied, one can show from the formulae for the absorption and emission coefficient that for an unbroken power law electron distribution with index $`p`$ the magnetic field strength may be expressed as $`B_{\mathrm{pl}}=2.488\times 10^{62}\left({\displaystyle \frac{a(p)}{b(p)}}\right)^2(1\mathrm{e}^{\tau _{\mathrm{max}}})`$ $`\left({\displaystyle \frac{4F_{\mathrm{max}}/(\pi \theta ^2)}{1\mathrm{erg}/\mathrm{cm}}}\right)^2\left({\displaystyle \frac{\nu _{\mathrm{max}}}{1\mathrm{Hz}}}\right)^5\mathrm{G}.`$ $`\tau _{\mathrm{max}}`$ constitutes the optical density corresponding to the frequency $`\nu _{\mathrm{max}}`$ of the maximum intensity. The functions $`a(p)`$ and $`b(p)`$ are of the order 1 and can be found in Longair (1994). The magnetic field actually present can be written as $$B=bB_{\mathrm{pl}},$$ where the factor $`b`$ is dependent on the electron spectrum (cf. Sections 3 and 4). In addition, the peak brightness temperature of the source is defined as $$T_{\mathrm{max}}\frac{c^2}{2k_\mathrm{B}\nu _{\mathrm{max}}^2}\frac{F_{\mathrm{max}}}{\pi \theta ^2/4}.$$ We now investigate the influence of the low energy electron spectrum on the limiting brightness temperature $`T_{\mathrm{max}}^\mathrm{c}10^{12}\mathrm{K}`$ for the onset of the inverse Compton catastrophe, i.e. the temperature $`T_{\mathrm{max}}`$ for which $`P_\mathrm{c}/P_{\mathrm{syn}}=1`$. Combining the formulae obtained so far, including second order inverse Compton losses, and choosing the power law index $`p=3`$ yields $`{\displaystyle \frac{P_\mathrm{c}}{P_{\mathrm{syn}}}}=0.52{\displaystyle \frac{\xi _{\mathrm{geom}}}{b^2}}\left({\displaystyle \frac{T_{\mathrm{max}}}{10^{12}\mathrm{K}}}\right)^5\left({\displaystyle \frac{\nu _\mathrm{c}}{1\mathrm{MHz}}}\right)`$ $`\left[1+0.52{\displaystyle \frac{\xi _{\mathrm{geom}}}{b^2}}\left({\displaystyle \frac{T_{\mathrm{max}}}{10^{12}\mathrm{K}}}\right)^5\left({\displaystyle \frac{\nu _\mathrm{c}}{1\mathrm{MHz}}}\right)\right].`$ If we assume that $`0.3<b=B/B_{\mathrm{pl}}<3`$, we see that the values of $`T_{\mathrm{max}}^\mathrm{c}`$ corresponding to $`B_{\mathrm{pl}}`$ and $`B`$ differ by a factor less than approximately $`3^{2/5}1.6`$. Hence, the uncertainty in the magnetic field determination due to the unknown optically thick electron spectrum has a rather small impact on the limiting brightness temperature. Indeed, the uncertainty in the source geometry (i.e. in $`\xi _{\mathrm{geom}}`$) may be considerably more important. ## 6 Conclusions Based on the discussion of stationary solutions to the kinetic equations in a homogeneous and isotropic synchrotron plasma, the influence of the optically thick part of the electron spectrum on the magnetic field determination from the position of the turnover in the photon spectrum has been investigated. By employing the assumption that the optical depth equals unity at the turnover frequency it was shown that the error made by choosing some specific form for the low energy electron spectrum can be estimated to be within one order of magnitude. For power law distributions this result was corroborated by a more rigorous treatment. All results obtained are independent from the spatial extension of the source and the overall number of electrons. They show, for example, that the limiting brightness temperature is influenced to a small extent only. Thus we have seen that that determining a magnetic field strength from the turnover of a synchrotron spectrum yields a value that is of the correct order of magnitude. Of course, in actual computations other uncertainties such as in the source geometry must also be taken into account. ## Acknowledgments We wish to thank Klaus Beuermann and Frank Rieger for many helpful discussions. Part of this work has been supported by the Graduiertenkolleg “Strömungsinstabilitäten und Turbulenz” of the Deutsche Forschungsgemeinschaft (DFG). ## Bibliography Ginzburg, V.L., Syrovatskii, S.I., 1965, ARA&A 3, 297 Jones, T.W., 1988, ApJ 332, 678 Kellermann, K.I., 1974, in Verschuur, G.L., Kellermann, K.I., eds, Galactic and Extra-Galactic Radio Astronomy. Springer, New York, p. 320 Longair, M.S., 1994, High energy astrophysics (Volume 2), Cambridge Univer- sity Press, Cambridge McCray, R., 1969, ApJ 156, 329 Melrose, D.B., 1980, Plasma Astrophysics (Volume 1), Gordon and Breach, New York Norman, C.A., ter Haar, D., 1974, Phys. Rep. 17, 307 Pacholczyk, A.G., 1970, Radio Astrophysics, W. H. Freeman and Company, San Francisco Rybicki, G.B., Lightman, A.P., 1979, Radiative Processes in Astrophysics,Wiley, New York Westfold, K.C., 1959, ApJ 130, 241
warning/0003/nlin0003036.html
ar5iv
text
# Hierarchy of Chaotic Maps with an Invariant Measure ## 1 INTRODUCTION In recent years chaos or more properly dynamical systems have become an important area of research activity. One of the landmarks in it was introduction of the concept of Sinai-Ruelle-Bowen (SRB) measure or natural invariant measure. This is roughly speaking a measure that is supported on an attractor and that describe the statistics of the long time behavior of the orbits for almost every initial condition in the corresponding basin of attractor . This measure can be obtained by computing the fixed density of the so called Frobenius-Perron operator which can be viewed as a differential-integral operator, hence, exact determination of invariant measure of dynamical systems is rather a nontrivial task, such that invariant measure of few dynamical systems such as one-parameter family one-dimensional piecewise linear maps including Baker and tent maps or unimodal maps such as logistic map for certain values of its parameter, can be derived analytically. In most of cases only numerical algorithms, as an example Ulam’s method are used for computation of fixed densities of Frobenius-Perron operator. Here in this article we give a hierarchy of one-parameter family $`\mathrm{\Phi }^{(1,2)}(\alpha ,x)`$ of maps of interval $`[0,1]`$ with an invariant measure. These maps are defined as ratio of polynomials of degree N, where we have derived analytically their invariant measure for an arbitrary values of the parameter $`\alpha `$ and every integer values of $`N`$. Using this measure, we have calculated analytically, Kolmogorov-Sinai entropy or equivalently positive Lyapunov characteristic exponent of these maps, where the numerical simulation of up to degree N=10 approve the analytic calculation. Also it is shown that these maps have another interesting property, that is, for even values of $`N`$ the $`\mathrm{\Phi }^{(1)}(\alpha ,x)`$($`\mathrm{\Phi }^{(2)}(\alpha ,x)`$) maps have only a fixed point attractor $`x=1`$($`x=0`$) provided that their parameter belongs to interval $`(N,\mathrm{})`$($`(0,\frac{1}{N})`$) while, at $`\alpha N`$ $`(\alpha \frac{1}{N})`$ they bifurcate to chaotic regime without having any period doubling or period-n-tupling scenario and remain chaotic for all $`\alpha (0,N)`$ ($`\alpha (\frac{1}{N},\mathrm{})`$) but for odd values of N, these maps have only fixed point attractor $`x=0`$ for $`\alpha (\frac{1}{N},N)`$, again they bifurcate to chaotic regime at $`\alpha \frac{1}{N}`$, and remain chaotic for $`\alpha (0,\frac{1}{N})`$, finally they bifurcate at $`\alpha =N`$ to have $`x=1`$ as fixed point attractor for all $`\alpha (\frac{1}{N},\mathrm{})`$(see Figures 1,2 and 3). The paper is organized as follows: In section II we introduce hierarchy of family of one-parameter maps, In Section III we show that the proposed anzats for the invariant measure of their maps are eigenfuntion of Ferobenios-Perron operator with largest eigenvalue $`1`$, for any finite $`N`$. Then in section IV using this measure we calculate kolmogorov-Sinai entropy of these maps for an arbitrary value of parameter $`\alpha `$ and every integer values of $`N`$. In section V we compare analytic calculation with the numerical simulation. Paper ends with a brief conclusion. ## 2 ONE-PARAMETER FAMILIES OF CHAOTIC MAPS The one-parameter families of chaotic maps of the interval $`[0,1]`$ with an invariant measure are defined as ratio of polynomials of degree $`N`$ : $$\mathrm{\Phi }_N^{(1)}(x,\alpha )=\frac{\alpha ^2\left(1+(1)^N{}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)\right)}{(\alpha ^2+1)+(\alpha ^21)(1)^N{}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)}$$ $$=\frac{\alpha ^2(T_N(\sqrt{x}))^2}{1+(\alpha ^21)(T_N(\sqrt{x})^2)},$$ (2-1) $$\mathrm{\Phi }_N^{(2)}(x,\alpha )=\frac{\alpha ^2\left(1(1)^N{}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},(1x))\right)}{(\beta ^2+1)(\alpha ^21)(1)^N{}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},(1x))}$$ $$=\frac{\alpha ^2(U_N(\sqrt{(1x)}))^2}{1+(\alpha ^21)(U_N(\sqrt{(1x)})^2)},$$ (2-2) where $`N`$ is an integer greater than one. Also $${}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)=(1)^N\mathrm{cos}(2N\mathrm{arccos}\sqrt{x})=(1)^NT_{2N}(\sqrt{x})$$ is hypergeometric polynomials of degree $`N`$ and $`T_N(x)(U_n(x))`$ are Chebyshev polynomials of type I (type II), respectively. Obviously these map the unit interval $`[0,1]`$ into itself and are related to each other through the following relation : $$\mathrm{\Phi }_N^{(1)}(x,\alpha )=g(\mathrm{\Phi }_N^{(2)},g(x),\frac{1}{\alpha })=g\mathrm{\Phi }_N^{(1)}(\frac{1}{\alpha })g(x)\text{for even N,}$$ (2-3) and $$\mathrm{\Phi }_N^{(1)}(x,\alpha )=\mathrm{\Phi }_N^{(2)}(x,\alpha )\text{for odd N,}$$ (2-4) where $`g(x)`$ is the invertible map $`g(x)=g^1(x)=1x`$ and the symbol $``$ means composition of functions. From now on, depending on situation, we will consider one of these maps, since, we can get all required information concerning the other map via using the relations $`(23)`$ and $`(24)`$ between these two maps: $`\mathrm{\Phi }_N^{(1)}(\alpha ,x)`$ is (N-1)-model map, that is it has $`(N1)`$ critical points in unit interval $`[0,1]`$,(see Figure 4) since its derivative is proportional to derivative of hypergeometric polynomial $`{}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)`$ which is itself a hypergeometric polynomial of degree $`(N1)`$, hence it has $`(N1)`$ real roots in unit interval $`[0,1]`$. Defining Shwarzian derivative $`S\mathrm{\Phi }_N(x)`$ as: $$S\left(\mathrm{\Phi }_N^{(1)}(x)\right)=\frac{\mathrm{\Phi }_N^{(1)\prime \prime \prime }(x)}{\mathrm{\Phi }_N^{(1)}(x)}\frac{3}{2}\left(\frac{\mathrm{\Phi }_N^{(1)\prime \prime }(x)}{\mathrm{\Phi }_N^{(1)}(x)}\right)^2=\left(\frac{\mathrm{\Phi }_N^{(1)\prime \prime }(x)}{\mathrm{\Phi }_N^{(1)}(x)}\right)^{}\frac{1}{2}\left(\frac{\mathrm{\Phi }_N^{(1)\prime \prime }(x)}{\mathrm{\Phi }_N^{(1)}(x)}\right)^2,$$ with a prime denoting a single differentiation with respect to variable $`x`$, one can show that: $$S\left(\mathrm{\Phi }_N^{(1)}(x)\right)=S\left({}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)\right)0,$$ since $`\frac{d}{dx}(_2F_1(N,N,\frac{1}{2},x))`$ can be written as: $$\frac{d}{dx}\left({}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)\right)=A\underset{i=1}{\overset{N1}{}}(xx_i),$$ with $`0x_1<x_2<x_3<\mathrm{}.<x_{N1}1`$ then we have: $$S\left({}_{2}{}^{}F_{1}^{}(N,N,\frac{1}{2},x)\right)=\frac{1}{2}\underset{J=1}{\overset{N1}{}}\frac{1}{(xa_j)^2}\left(\underset{J=1}{\overset{N1}{}}\frac{1}{(xx_j)}\right)^2<0.$$ Therefore the maps $`\mathrm{\Phi }_N^{(1)}(x)`$ have at most $`N+1`$ attracting periodic orbits. As we will show at the end of this section, these maps have only a single period one stable fixed points. Denoting n-composition of functions $`\mathrm{\Phi }^{(1,2)}(x_1,\alpha )`$ by $`\mathrm{\Phi }^{(n)}`$, it is straightforward to show that the derivative of $`\mathrm{\Phi }^{(n)}`$ at its possible n periodic points of an n-cycle: $`x_2=\mathrm{\Phi }^{(1,2)}(x_1,\alpha ),x_3=\mathrm{\Phi }^{(1,2)}(x_2,\alpha ),\mathrm{},x_1=\mathrm{\Phi }^{(1,2)}(x_n,\alpha )`$ is $$\frac{d}{dx}\mathrm{\Phi }^{(n)}=\frac{d}{dx}\stackrel{n}{\stackrel{}{\left(\mathrm{\Phi }^{(1,2)}\mathrm{\Phi }^{(1,2)}\mathrm{}\mathrm{\Phi }^{(1,2)}(x,\alpha )\right)}}=\underset{k=1}{\overset{n}{}}\frac{N}{\alpha }(\alpha ^2+(1\alpha ^2)x_k),$$ (2-5) since for $`x_k[0,1]`$ we have: $$min(\alpha ^2+(1\alpha ^2x_k))=min(1,\alpha ^2),$$ therefore, $$min\frac{d}{dx}\mathrm{\Phi }^{(n)}=\left(\frac{N}{\alpha }min(1,\alpha ^2)\right)^n.$$ Hence the above expression is definitely greater than one for $`\frac{1}{N}<\alpha <N`$, that is, both maps do not have any kind of n-cycle or periodic orbits for $`\frac{1}{N}<\alpha <N`$, actually they are ergodic for this interval of parameter. From (2-13) it follows that $`\frac{d}{dx}\mathrm{\Phi }^{(n)}`$ at n periodic points of the n-cycle belonging to interval , varies between $`(N\alpha )^n`$ and $`(\frac{N}{\alpha })^n`$ for $`\alpha <\frac{1}{N}`$ and between $`(\frac{N}{\alpha })^n`$ and $`(N\alpha )^n`$ for $`\alpha >N`$, respectively. From the definition of these maps, we see that for odd N, both $`x=0`$ and $`x=1`$ belong to one of the n-cycles, while for even N, only $`x=1`$ belongs to one of the n-cycles of $`\mathrm{\Phi }_N^{(1)}(x,\alpha )`$ and $`x=0`$ belongs to one of the n-cycles of $`\mathrm{\Phi }_N^{(2)}(x,\alpha ).`$ For $`\alpha <(\frac{1}{N})(\alpha >N)`$, the formula $`(25)`$ implies that for those cases in which $`x=0(x=1)`$ belongs to one of n-cycles we will have $`\frac{d}{dx}\mathrm{\Phi }^{(n)}<1`$, hence the curve of $`\mathrm{\Phi }^{(n)}`$ starts at $`x=0(x=1)`$ beneath the bisector and then crosses it at the next (previous) periodic point with slope greater than one (see Fig. 1), since the formula $`(25)`$ implies that the slope of fixed points increases with the increasing (decreasing) of $`x_k`$, therefore at all periodic points of n-cycles except for $`x=0(x=1)`$ the slop is greater than one that is they are unstable, this is possible only if $`x=0(x=1)`$ is the only period one fixed point of these maps. Hence all n-cycles except for possible period one fixed points $`x=0`$ and $`x=1`$ are unstable, where for $`\alpha [0,\frac{1}{N}]`$, the fixed point $`x=0`$ is stable in maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$(for odd N) and $`\mathrm{\Phi }_N^{(2)}(x,\alpha )`$ (for even N), while for $`\alpha [N,\mathrm{})`$ and $`\mathrm{\Phi }_N^{(1)}(x,\alpha )`$, the $`x=1`$ is stable fixed point in maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$(for odd N). As an example we give below some of these maps: $$\varphi _2^{(1)}=\frac{\alpha ^2(2x1)^2}{4x(1x)+\alpha ^2(2x1)^2},$$ $$\varphi _2^{(2)}=\frac{4\alpha ^2x(1x)}{1+4(\alpha ^21)x(1x)},$$ $$\varphi _3^{(1)}=\varphi _3^{(2)}=\frac{\alpha ^2x(4x3)^2}{\alpha ^2x(4x3)^2+(1x)(4x1)^2},$$ $$\varphi _4^{(1)}=\frac{\alpha ^2(18x(1x))^2}{\alpha ^2(18x(1x))^2+16x(1x)(12x)^2},$$ $$\varphi _4^{(2)}=\frac{16\alpha ^2x(1x)(12x)^2}{(18x+8x^2)^2+16\alpha ^2x(1x)(12x)^2},$$ $$\varphi _5^{(1)}=\varphi _5^{(2)}=\frac{\alpha ^2x(16x^220x+5)^2}{\alpha ^2x(16x^220x+5)^2+(1x)(16x^2(2x1))}.$$ Below we also introduce their conjugate or isomorphic maps which will be very useful in derivation of their invariant measure and calculation of their KS-entropy in the next section. Conjugacy means that the invertible map $`h(x)=\frac{1x}{x}`$ maps $`I=[0,1]`$ into $`[0,\mathrm{})`$ and transforms maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ into $`\stackrel{~}{\mathrm{\Phi }}_N^{(1,2)}(x,\alpha )`$ defined as: $$\{\begin{array}{c}\stackrel{~}{\mathrm{\Phi }}_N^{(1)}(x,\alpha )=h\mathrm{\Phi }_N^{(1)}(x,\alpha )h^{(1)}=\frac{1}{\alpha ^2}\mathrm{tan}^2(N\mathrm{arctan}\sqrt{x}),\hfill \\ \stackrel{~}{\mathrm{\Phi }}_N^{(2)}(x,\alpha )=h\mathrm{\Phi }_N^{(2)}(x,\alpha )h^1=\frac{1}{\alpha ^2}\mathrm{cot}^2(N\mathrm{arctan}\frac{1}{\sqrt{x}}).\hfill \end{array}$$ (2-6) ## 3 INVARIANT MEASURE Dynamical systems, even apparently simple dynamical systems as those described by maps of an interval can display a rich variety of different asymptotic behavior. On measure theoretical level these types of behavior are described by SRB or invariant measure describing statistically stationary states of the system. The probability measure $`\mu `$ on $`[0,1]`$ is called an SRB or invariant measure of the maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ given in $`(21)`$ and $`(22)`$, if it is $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$-Invariant and absolutely continuous with respect to Lebesgue measure. For deterministic system such as $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$-map, the $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$-invariance means that its invariant measure $`\mu (x)`$ fulfills the following formal Ferbenius-Perron integral equation $$\mu (y)=_0^1\delta (y\mathrm{\Phi }_N^{(1,2)}(x,\alpha ))\mu (x)𝑑x.$$ This is equivalent to: $$\mu (y)=\underset{x\mathrm{\Phi }_N^{1(1,2)}(y,\alpha )}{}\mu (x)\frac{dx}{dy},$$ (3-1) defining the action of standard Ferobenius-Perron operator for the map $`\mathrm{\Phi }_N(x)`$ over a function as: $$P_{\mathrm{\Phi }_N^{(1,2)}}f(y)=\underset{x\mathrm{\Phi }_N^{1(1,2)}(y,\alpha )}{}f(x)\frac{dx}{dy}.$$ (3-2) We see that, the invariant measure $`\mu (x)`$ is given as the eigenstate of the Frobenios-Perron operator $`P_{\mathrm{\Phi }_N^{(1,2)}}`$ corresponding to largest eigenvalue 1. As we will prove below the measure $`\mu _{\mathrm{\Phi }_N^{(1,2)}(x,\alpha )}(x,\beta )`$ defined as: $$\frac{1}{\pi }\frac{\sqrt{\beta }}{\sqrt{x(1x)}(\beta +(1\beta )x)},$$ (3-3) with $`\beta >0`$ is the invariant measure of the maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ provided that we choose the parameter $`\alpha `$ in the following form : $$\alpha =\frac{{}_{k=0}{}^{[\frac{(N1)}{2}]}C_{2k+1}^{N}\beta ^k}{{}_{k=0}{}^{[\frac{N}{2}]}C_{2k}^{N}\beta ^k},$$ (3-4) in $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ maps for odd values of N and in $`\mathrm{\Phi }_N^{(1)}(x,\alpha )`$maps for even values of N $$\alpha =\frac{\beta {}_{k=0}{}^{[\frac{(N)}{2}]}C_{2k}^{N}\beta ^k}{{}_{k=0}{}^{[\frac{(N1)}{2}]}C_{2k+1}^{N}\beta ^k},$$ (3-5) in $`\mathrm{\Phi }_N^{(2)}(x,\alpha )`$ maps for even values of N, where $`[]`$ means greatest integer part. As we see the above measure is defined only for $`\beta >0`$ hence, from the relations $`(34)`$ and $`(35)`$, it follows that the maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ have invariant measure only for $`\alpha (\frac{1}{N},N)`$ for odd values of N, $`\mathrm{\Phi }_N^{(1)}(x,\alpha )`$ maps have invariant measure for $`\alpha (0,N)`$ and $`\mathrm{\Phi }_N^{(2)}(x,\alpha )`$ have for $`\alpha (\frac{1}{N},\mathrm{})`$ for even N, respectively. For other values of $`\alpha `$ these maps have single attractive fixed points, which is the same as the prediction of the previous section. In order to prove that measure $`(33)`$ satisfies equation $`(31)`$, with $`\alpha `$ given by relations $`(34)`$ and $`(35)`$, it is rather convenient to consider the conjugate map: $$\stackrel{~}{\mathrm{\Phi }}_N^{(1)}(x,\alpha )=\frac{1}{\alpha ^2}\mathrm{tan}^2(N\mathrm{arctan}\sqrt{x}),$$ (3-6) with measure $`\stackrel{~}{\mu }_{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}`$related to the measure $`\mu _{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}`$ to the following relation: $$\stackrel{~}{\mu }_{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}(x)=\frac{1}{(1+x)^2}\mu _{\mathrm{\Phi }_N^{(1)}}(\frac{1}{1+x}).$$ Denoting $`\stackrel{~}{\mathrm{\Phi }}_N^{(1)}(x,\alpha )`$ on the left hand side of $`(36)`$ by $`y`$ and inverting it, we get : $$x_k=\mathrm{tan}^2(\frac{1}{N}\mathrm{arctan}\sqrt{y\alpha ^2}+\frac{k\pi }{N})k=1,..,N.$$ (3-7) Then, taking derivative of $`x_k`$ with respect to $`y`$, we obtain: $$\frac{dx_k}{dy}=\frac{\alpha }{N}\sqrt{x_k(1+x_k)}\frac{1}{\sqrt{y}(1+\alpha ^2y)}.$$ (3-8) Substituting the above result in equation $`(31)`$, we have: $$\stackrel{~}{\mu }_{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}(y)\sqrt{y}(1+\alpha ^2y)=\frac{1}{N}\underset{k}{}\sqrt{x_k}(1+x_k\stackrel{~}{\mu }_{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}(x_k)),$$ (3-9) considering the following anzats for the invariant measure $`\stackrel{~}{\mu }_{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}(y)`$: $$\stackrel{~}{\mu }_{\stackrel{~}{\mathrm{\Phi }}_N^{(1)}}(y)=\frac{1}{\sqrt{y}(1+\beta y)},$$ (3-10) the above equation reduces to: $$\frac{1+\alpha ^2y}{1+\beta y}=\frac{\alpha }{N}\underset{k=1}{\overset{N}{}}\left(\frac{1+x_k}{1+\beta x_k}\right)$$ which can be written as: $$\frac{1+\alpha ^2y}{1+\beta y}=\frac{\alpha }{\beta }+\left(\frac{\beta 1}{\beta ^2}\right)\frac{}{\beta ^1}(\mathrm{ln}(\mathrm{\Pi }_{k=1}^N(\beta ^1+x_k))).$$ (3-11) To evaluate the second term in the right hand side of above formulas we can write the equation in the following form: $$0=\alpha ^2y\mathrm{cos}^2(N\mathrm{arctan}\sqrt{x})\mathrm{sin}^2(N\mathrm{arctan}\sqrt{x})$$ $$=\frac{(1)^N}{(1+x)^N}\left(\alpha ^2y(\underset{k=0}{\overset{[\frac{N}{2}]}{}}C_{2k}^N(1)^Nx^k)^2x(\underset{k=0}{\overset{[\frac{N1}{2}]}{}}C_{2k+1}^N(1)^Nx^k)^2\right),$$ $$=\frac{\text{constant}}{(1+x)^N}\underset{k=1}{\overset{N}{}}(xx_k),$$ where $`x_k`$ are the roots of equation $`(36)`$, and are given by formula $`(37)`$.Therefore, we have: $$\frac{}{\beta ^1}\mathrm{ln}\left(\underset{k=1}{\overset{N}{}}(\beta ^1+x_k)\right)$$ $$=\frac{}{\beta ^1}\mathrm{ln}\left[(1\beta ^1)^N(\alpha ^2y\mathrm{cos}^2(N\mathrm{arctan}\sqrt{\beta ^1})\mathrm{sin}^2(N\mathrm{arctan}\sqrt{\beta ^1}))\right]$$ $$=\frac{N\beta }{\beta 1}+\frac{\beta N(1+\alpha ^2y)A(\frac{1}{\beta })}{(A(\frac{1}{\beta }))^2\beta ^2y+(B(\frac{1}{\beta }))^2},$$ (3-12) with polynomials $`A(x)`$ and $`B(x)`$ defined as: $$A(x)=\mathrm{\Pi }_{k=0}^{[\frac{N}{2}]}C_{2k}^Nx^k,$$ $$B(x)=\mathrm{\Pi }_{k=0}^{[\frac{N1}{2}]}C_{2k+1}^Nx^k.$$ (3-13) In deriving the of above formula we have used the following identities: $$\mathrm{cos}(N\mathrm{arctan}\sqrt{x})=\frac{A(x)}{(1+x)^{\frac{N}{2}}},$$ $$\mathrm{sin}(N\mathrm{arctan}\sqrt{x})=\sqrt{x}\frac{B(x)}{(1+x)^{\frac{N}{2}}},$$ (3-14) inserting the results $`(312)`$, in $`(36)`$, we get: $$\frac{1+\alpha ^2y}{1+\beta y}=\frac{1+\alpha ^2y}{(\frac{B(\frac{1}{\beta })}{\alpha A(\frac{1}{\beta })}+\beta (\frac{\alpha A(\frac{1}{\beta })}{B(\frac{1}{\beta })})}.$$ Hence to get the final result we have to choose the parameter $`\alpha `$ as: $$\alpha =\frac{B(\frac{1}{\beta })}{A(\frac{1}{\beta })}.$$ With the procedure similar to the one given above we could get the relation $`(311)`$ between the parameters $`\alpha `$ and $`\beta `$ for the second kind of maps. ## 4 KOLMOGROV-SINAI ENTROPY Kolomogrov-Sinai entropy (KS) or metric entropy measure how chaotic a dynamical system is and it is proportional to the rate at which information about the state of dynamical system is lost in the course of time or iteration. Therefore, it can also be defined as the average rate of loss information for a discrete measurable dynamical system $`(\mathrm{\Phi }_N^{(1,2)}(x,\alpha ),\mu )`$, by introducing a partition $`\alpha =A_c(n_1,\mathrm{}..n_\gamma )`$ of the interval $`[0,1]`$ into individual laps $`A_i`$ one can define the usual entropy associated with the partition by: $$H(\mu ,\gamma )=\underset{i=1}{\overset{n(\gamma )}{}}m(A_c)\mathrm{ln}m(A_c),$$ where $`m(A_c)={}_{nA_i}{}^{}\mu (x)𝑑x`$ is the invariant measure of $`A_i`$. Defining n-th refining $`\gamma (n)`$ of $`\gamma `$: $$\gamma ^n=\underset{k=0}{\overset{n1}{}}(\mathrm{\Phi }_N^{(1,2)}(x,\alpha ))^{(k)}(\gamma ),$$ and defining an entropy per unit step of refining by : $$h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha ),\gamma )=lim{}_{n\mathrm{}}{}^{}(\frac{1}{n}H(\mu ,\gamma )),$$ if the size of individual laps of $`\gamma (N)`$ tends to zero as n increases, then the above entropy is known as Kolmogorov-Sinai entropy, that is: $$h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha ))=h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha ),\gamma ).$$ KS-entropy , which is a quantitative measure of the rate of information lost with the refining, may also be written as: $$h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha ))=\mu (x)𝑑x\mathrm{ln}\frac{d}{dx}\mathrm{\Phi }_N^{(1,2)}(x,\alpha ),$$ (4-1) which is also a statistical mechanical expression for the Lyapunov characteristic exponent, that is, mean divergence rate of two nearby orbits. The measurable dynamical system $`(\mathrm{\Phi }_N^{(1,2)}(x,\alpha ),\mu )`$ is chaotic for $`h>0`$ and predictive for $`h=0`$. In order to calculate the KS-entropy of the maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$, it is rather convenient to consider their conjugate maps given by $`(26)`$, since it can be shown that KS-entropy is a kind of topological invariant, that is, it is preserved under conjugacy map, hence we have: $$h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha ))=h(\stackrel{~}{\mu },\stackrel{~}{\mathrm{\Phi }}_N^{(1,2)}(x,\alpha )).$$ (4-2) Using the integral $`(41)`$, the KS-entropy of $`\mathrm{\Phi }_N^{(2)}(\alpha ,x)`$ can be writen as $$h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=h(\stackrel{~}{\mu },\stackrel{~}{\mathrm{\Phi }}_N^{(2)}(x,\alpha ))$$ $$=\frac{1}{\pi }_0^{\mathrm{}}\frac{\sqrt{\beta }dx}{\sqrt{x}(1+\beta x)}\mathrm{ln}(\frac{1}{a^2}\frac{d}{dx}(\mathrm{cot}^2(N\mathrm{arctan}\sqrt{x})))$$ $$=\frac{1}{\pi }_0^{\mathrm{}}\frac{\beta dx}{\sqrt{x}(1+\beta x)}ln\left(\frac{N}{\alpha ^2}\times \frac{1}{\sqrt{x}(1+x)}\frac{\mathrm{cos}N(\mathrm{arctan}\sqrt{x})}{\mathrm{sin}^3N(\mathrm{arctan}\sqrt{x})}\right).$$ (4-3) Using the relations given in $`(314)`$ we have $$h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\frac{1}{\pi }_0^{\mathrm{}}\frac{\sqrt{\beta }dx}{\sqrt{x}(1+\beta x)}\mathrm{ln}\left(\frac{N}{\alpha ^2}\frac{(1+x)^{N1}A(x)}{x^2(B(x))^3}\right),$$ (4-4) for even N, and $$h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\frac{1}{\pi }_0^{\mathrm{}}\frac{\sqrt{\beta }dx}{\sqrt{x}(1+\beta x)}\mathrm{ln}\left(\frac{N}{\alpha ^2}\frac{(1+x)^{N1}B(x)}{(A(x))^3}\right).$$ (4-5) for odd N. Considering again the relations given in $`(314)`$, we see that polynomials appearing in the numerator ( denominator ) of integrand appearing on the right hand side of equation $`(45)`$, have $`\frac{[N1]}{2}`$ $`(\frac{[N]}{2})`$ simple roots, denoted by $`x_k^Bk=1,\mathrm{},[\frac{N1}{2}]`$ $`(x_k^Ak=1,\mathrm{},[\frac{N}{2}])`$ in the interval $`[0,\mathrm{})`$. Hence, we can write the above formula in the following form: $$h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\frac{1}{\pi }_0^{\mathrm{}}\frac{\sqrt{\beta }dx}{\sqrt{x}(1+\beta x)}\mathrm{ln}\left(\frac{N}{\alpha ^2}\times \frac{(1+x)^{N1}_{k=1}^{[\frac{N}{2}]}xx_k^A}{x^2_{k=1}^{[\frac{N1}{2}]}xx_k^B^3}\right),$$ for even N, and $$h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\frac{1}{\pi }_0^{\mathrm{}}\frac{\sqrt{\beta }dx}{\sqrt{x}(1+\beta x)}\mathrm{ln}\left(\frac{N}{\alpha ^2}\times \frac{(1+x)^{N1}_{k=1}^{[\frac{N1}{2}]}xx_k^B}{_{k=1}^{[\frac{N}{2}]}xx_k^A^3}\right).$$ for odd N. Now making the following change of variable $`x=\frac{1}{\beta }\mathrm{tan}^2\frac{\mathrm{\Theta }}{2}`$, and taking into account that degree of numerators and denominator are equal for both even and odd values of N, we get $$h(\alpha ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\frac{1}{\pi }_0^{\mathrm{}}d\theta \{\mathrm{ln}(\frac{N}{\alpha ^2})+(N1)\mathrm{ln}(\beta +1+(\beta 1)\mathrm{cos}\theta )$$ $$+\underset{k=1}{\overset{[\frac{N}{2}]}{}}\mathrm{ln}1x_k^A\beta +(1+x_k^A\beta )\mathrm{cos}\theta 3\underset{k=1}{\overset{[\frac{N1}{2}]}{}}\mathrm{ln}1x_k^B\beta +(1+x_k^B\beta )\mathrm{cos}\theta $$ $$2\mathrm{ln}1+cos\theta \},$$ for even and $$h(\alpha ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\frac{1}{\pi }_0^{\mathrm{}}d\theta \{\mathrm{ln}(\frac{N}{\alpha ^2})+(N1)\mathrm{ln}(\beta +1+(\beta 1)\mathrm{cos}\theta )$$ $$+\underset{k=1}{\overset{[\frac{N1}{2}]}{}}\mathrm{ln}1x_k^B\beta +(1+x_k^B\beta )\mathrm{cos}\theta 3\underset{k=1}{\overset{[\frac{N}{2}]}{}}\mathrm{ln}1x_k^A\beta +(1+x_k^A\beta )\mathrm{cos}\theta \}.$$ for odd N. Using the following integrals: $$\frac{1}{\pi }_0^\pi \mathrm{ln}a+b\mathrm{cos}\theta =\{\begin{array}{c}\mathrm{ln}\frac{a+\sqrt{a^2b^2}}{2}a>b\hfill \\ \mathrm{ln}\frac{b}{2}ab,\hfill \end{array}$$ we get $$h(\alpha ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\{\begin{array}{c}\mathrm{ln}\left(\frac{N}{\alpha ^2}\frac{(\beta +1+2\sqrt{\beta })^{N1}_{k=1}^{[\frac{N}{2}]}(1+x_k^A\beta )}{\left(_{k=1}^{[\frac{N1}{2}]}(1+x_k^B\beta )\right)^3}\right)\text{for even N}\hfill \\ \\ \mathrm{ln}\left(\frac{N}{\alpha ^2}\frac{(\beta +1+2\sqrt{\beta })^{N1}_{k=1}^{[\frac{N1}{2}]}(1+x_k^B\beta )}{\left(_{k=1}^{[\frac{N}{2}]}(1+x_k^A\beta )\right)^3}\right)\text{for odd N},\hfill \end{array}$$ or $$h(\alpha ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\{\begin{array}{c}\mathrm{ln}\left(\frac{N}{\alpha ^2}\frac{(1+\beta +2\sqrt{\beta })^{N1}\beta ^2A(\frac{1}{\beta })}{\beta ^{(N1)}\left(B(\frac{1}{\beta })\right)^3}\right)\text{for even N}\hfill \\ \\ \mathrm{ln}\left(\frac{N}{\alpha ^2}\frac{(1+\beta +2\sqrt{\beta })^{N1}B(\frac{1}{\beta })}{\beta ^{(N1)}\left(A(\frac{1}{\beta })\right)^3}\right)\text{for odd N}.\hfill \end{array}$$ Using the relation: $$\alpha =\{\begin{array}{c}\beta \frac{A(\frac{1}{\beta })}{B(\frac{1}{\beta })}\text{for even N}\hfill \\ \\ \frac{B(\frac{1}{\beta })}{A(\frac{1}{\beta })}\text{for odd N},\hfill \end{array}$$ we get $$h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha ))=\mathrm{ln}\left(\frac{N(1+\beta +2\sqrt{\beta })^{N1}}{(\mathrm{\Pi }_{k=0}^{[\frac{N}{2}]}C_{2k}^N\beta ^k)(\mathrm{\Pi }_{k=0}^{[\frac{N1}{2}]}C_{2k+1}^N\beta ^k)}\right).$$ (4-6) With a calculation rather similar to the one given above we can calculate the KS-entropy of the $`\mathrm{\Phi }_N^1(x,\alpha )`$ maps where the results are the same as those given by $`(46)`$. The KS-entropy $`(46)`$ is invariant with respect to $`\beta (\frac{1}{\beta })`$, therefore, it has the same asymptotic behavior near $`\beta 0`$ and $`\beta \mathrm{}`$ where its asymptotic forms read $$\{\begin{array}{c}h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha =N+0^{}))(N\alpha )^{\frac{1}{2}}\hfill \\ h(\mu ,\mathrm{\Phi }_N^{(1,2)}(x,\alpha =\frac{1}{N}+0^+))(\alpha \frac{1}{N})^{\frac{1}{2}},\hfill \end{array}$$ for odd N and: $$\{\begin{array}{c}h(\mu ,\mathrm{\Phi }_N^{(1)}(x,\alpha =N+0^{}))(N\alpha )^{\frac{1}{2}}\hfill \\ h(\mu ,\mathrm{\Phi }_N^{(2)}(x,\alpha =\frac{1}{N}+0^+))(\alpha \frac{1}{N})^{\frac{1}{2}},\hfill \end{array}$$ for even N. The above asymptotic form indicates that the maps, $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ belong to the same universality class which is different from the universality class of pitch fork bifurcating maps but their asymptotic behavior is similar to class of intermittent maps, even though intermittency can not occur in these maps for any values of parameter $`\alpha `$, since the maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ and their n-composition $`\mathrm{\Phi }^{(n)}`$ do not have minimum values other than zero and maximum values other than one in $`[0,1].`$ ## 5 SIMULATION Here in this section we try to calculate Lyapunov characteristic exponent of maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$, $`N=1,2,\mathrm{}.,5`$ in order to investigate these maps numerically. In fact, Lyapunov characteristic exponent is the characteristic exponent of the rate of average magnificent of the neighborhood of an arbitrary point $`x_0`$ and it is denoted by $`\mathrm{\Lambda }(x_0)`$ which is written as: $$\mathrm{\Lambda }^{(1,2)}(x_0)=lim_n\mathrm{}\mathrm{ln}(\stackrel{n}{\stackrel{}{\mathrm{\Phi }_N^{(1,2)}(x,\alpha )\mathrm{\Phi }_N^{(1,2)}\mathrm{}.\mathrm{\Phi }_N^{(1,2)}(x_K,\alpha )}}$$ $$=lim_n\mathrm{}\underset{k=0}{\overset{n1}{}}\mathrm{ln}\frac{d\mathrm{\Phi }_N^{(1,2)}(x_k,\alpha )}{dx},$$ (5-1) where $`x_k=\stackrel{k}{\stackrel{}{\mathrm{\Phi }_N^{(1,2)}\mathrm{\Phi }_N^{(1,2)}\mathrm{}.\mathrm{\Phi }_N^{(1,2)}(x_0)}}`$ . It is obvious that $`\mathrm{\Lambda }^{(1,2)}(x_0)<0`$ for an attractor, $`\mathrm{\Lambda }^{(1,2)}(x_0)>0`$ for a repeller and $`\mathrm{\Lambda }^{(1,2)}(x_0)=0`$ for marginal situation. Also the Liapunov number is independent of initial point $`x_0`$, provided that the motion inside the invariant manifold is ergodic, thus $`\mathrm{\Lambda }^{(1,2)}(x_0)`$ characterizes the invariant manifold of $`\mathrm{\Phi }_N^{(1,2)}`$ as a whole. For values of parameter $`\alpha `$ or $`\beta `$, such that the map $`\mathrm{\Phi }_N^{(1,2)}`$ be measurable, Birkohf ergodic theorem implies equality of KS-entropy and Liapunov characteristic exponent, that is: $$h(\mu ,\mathrm{\Phi }_n^{(1,2)})=\mathrm{\Lambda }^{(1,2)}(x_0,\mathrm{\Phi }_N^{(1,2)}),$$ (5-2) Comparison of analytically calculated KS-entropy of maps $`\mathrm{\Phi }_N^{(1,2)}(x,\alpha )`$ for $`N=1,2,\mathrm{}10`$ , (see Figures $`5,6`$ and $`7`$ for $`N=2`$ and $`3`$ ) with the corresponding Lyapunov characteristic exponent obtained by simulation, indicate that in chaotic region, these maps are ergodic as Birkohf ergodic Theorem predicts. In non chaotic region of parameter Lyapunov characteristic exponent is negative, since in this region we have only stable period one fixed points without bifurcation. In summation, combining the analytic discussion of section II with the numerical simulation we deduce that these maps are ergodic in certain values of their parameter as explained above and in complementary interval of parameter they have only a single period one attractive fixed point, such that in contrary to the most of usual one-dimensional one-parameter family of maps they have only a bifurcation from a period one attractive fixed point to chaotic state or vise versa. ## 6 Conclusion We have given hierarchy of exactly solvable one-parameter family of one-dimensional chaotic maps with an invariant measure, that is measurable dynamical system with an interesting property of being either chaotic (proper to say ergodic ) or having stable period one fixed point and they bifurcate from a stable single periodic state to chaotic one and vice-versa without having usual period doubling or period-n-tupling scenario. Perhaps this interesting property is due to existence of invariant measure for a range of values of parameter of these maps. Hence, to approve this conjecture, it would be interesting to find other measurable one parameter maps, specially higher dimensional maps, which is under investigation.
warning/0003/quant-ph0003116.html
ar5iv
text
# Physical implementation for entanglement purification of Gaussian continuous variable quantum states ## I Introduction Quantum entanglement plays an essential role in many interesting quantum information protocols, such as in quantum key distribution and quantum teleportation . To faithfully realize these protocols, first we need to generate a maximally entangled state. In reality, however, due to loss and decoherence, normally we can only generate partially entangled states between distant sides . Entanglement purification is further needed which distills a maximally entangled state from several pairs of partially entangled states using local quantum operations and classical communications . For qubit systems, efficient entanglement purification protocols have been found . Recently, quantum information protocols have been extended from qubit systems to continuous variable systems, such as continuous variable teleportation , continuous variable computation and error correction , continuous variable cryptography , and also the notions of continuous variable inseparability and bound entanglement have been investigated. For physical implementation, Gaussian continuous variable entangled states (i.e., states whose Wigner functions are Gaussians) can be generated experimentally by transmitting two-mode squeezed light, and this kind of entanglement has been demonstrated in the recent experiment of continuous variable teleportation . Obviously, it is useful to consider purification of continuous variable entanglement, that is, to generate a desired more entangled state from some realistic continuous entangled states. We have recently proposed an efficient continuous variable entanglement purification protocol . In this paper, we present the mathematical details of this purification protocol together with new results on its physical implementation. In particular, we take into account many important imperfections in a realistic experimental setup, and calculate their influence on the purification scheme. Quantitative requirements are given for the relevant experimental parameters. These calculations make necessary preparations for a real experiment. We also show how to generate Gaussian continuous entangled states between two distant high finesse cavities, which is the first step for the physical implementation of the purification protocol. It should be noted that with direct extensions of the purification protocols for qubit systems, it is possible to increase entanglement for a special class of less realistic continuous entangled states . Unfortunately, with these direct extensions no entanglement increase has been found till now for realistic Gaussian continuous entangled states. In a protocol to increase the entanglement for the special case of pure two-mode squeezed states has been proposed, which is based on conditional photon subtraction. For its practical realization, the efficiency, however, seems to be an issue. In contrast, the purification scheme discussed in this paper has the following favorable properties: (i) For pure states it reaches the maximal allowed efficiency in the asymptotic limit (when the number of pairs of modes goes to infinity); (ii) It can be readily extended to distill maximally entangled states from a relevant class of mixed Gaussian states which result from losses in the light transmission; (iii) An experimental scheme is possible for physical implementation of the purification protocol using high finesse cavities and cross Kerr nonlinearities. The paper is arranged as follows: In section 2 we show how to generate a Gaussian continuous entangled state between two distant cavities from the broadband squeezed light provided by a nondegenerate optical parametric amplifier (NOPA). Light transmission loss is taken into account. In sections 3 and 4 we give a detailed description of the purification protocol. Section 3 shows how to generate a maximally entangled state from pure two-mode squeezed states based on a local quantum non-demolition (QND) measurement of the total photon number, and section 4 extends the purification protocol to include the mixed Gaussian continuous states which are evolved from the pure two-mode squeezed states due to the unavoidable light transmission loss. In section 5, we describe a cavity scheme to realize the local QND measurement of the total photon number, and deduce conditions for the QND measurement. Then, in section 6, we extensively discuss many imperfections for a real experiment on QND measurements, and deduce quantitative requirements for the relevant experimental parameters. Last, we summarize the results, and give some typical parameter estimations. ## II Generation of continuous entangled states between two distant cavities Our source of entangled light field is taken to be a NOPA operating below threshold . The light fields may be nondegenerate in polarization or in frequency. The two NOPA cavity modes $`c_A`$ and $`c_B`$ are assumed to have the same output coupling rate $`\kappa _c`$. The dynamic in the NOPA cavity is described by the Langevin equations (in the rotating frame) $`\underset{A}{\overset{.}{c}}`$ $`=`$ $`ϵc_B^{}{\displaystyle \frac{\kappa _c}{2}}c_A\sqrt{\kappa _c}c_{iA},`$ (1) $`\underset{B}{\overset{}{\stackrel{.}{c}}}`$ $`=`$ $`ϵ^{}c_A{\displaystyle \frac{\kappa _c}{2}}c_B^{}\sqrt{\kappa _c}c_{iB}^{},`$ () where $`ϵ`$ is the pumping rate with $`\left|ϵ\right|<\kappa _c/2`$ (below threshold), and $`c_{iA}`$ and $`c_{iB}`$ are vacuum inputs. The NOPA outputs $`c_{oA}`$ and $`c_{oB}`$ are given respectively by $`c_{o\alpha }=c_{i\alpha }+\sqrt{\kappa _c}c_\alpha `$ $`\left(\alpha =A,B\right)`$. The two outputs, perhaps after a long distance propagation, are incident on distant high finesse cavities A and B. The cavities A and B are assumed to have the same damping rate $`\kappa `$ with $`\kappa \kappa _c`$. The schematic setup is shown by Fig. 1. Under the condition $`\kappa \kappa _c`$, the dynamics in the NOPA cavity is much faster than those in the cavities A and B, so we can assume a steady state for the NOPA outputs. The steady NOPA outputs are described by squeezed white noise operators with the following correlations $`c_{oA}\left(t\right)c_{oB}\left(t^{}\right)`$ $`=`$ $`M\delta \left(tt^{}\right),`$ (2) $`c_{o\alpha }^{}\left(t\right)c_{o\alpha }\left(t^{}\right)`$ $`=`$ $`N\delta \left(tt^{}\right),\text{ }\left(\alpha =A,B\right),`$ () $`c_{o\alpha }\left(t\right)c_{o\alpha }^{}\left(t^{}\right)`$ $`=`$ $`\left(N+1\right)\delta \left(tt^{}\right),\text{ }\left(\alpha =A,B\right),`$ (3) where $`N`$ and $`M`$, satisfying $`M=\sqrt{N\left(N+1\right)}`$, are determined by the NOPA coupling and pumping rates through $`N=\left|ϵ\right|^2\kappa _c^2/\left(\frac{\kappa _c^2}{4}\left|ϵ\right|^2\right)^2`$ and $`M=\left|ϵ\right|\kappa _c\left(\frac{\kappa _c^2}{4}+\left|ϵ\right|^2\right)/\left(\frac{\kappa _c^2}{4}\left|ϵ\right|^2\right)^2`$. To get the steady state of the cavities A and B, we note that their inputs $`a_{iA}`$ and $`a_{iB}`$ are respectively the NOPA outputs $`c_{oA}`$ and $`c_{oB}`$ with neglect of the losses during light propagation. The Langevin equations for the cavity modes $`a_A`$ and $`a_B`$ have the form $`\underset{\alpha }{\overset{.}{a}}={\displaystyle \frac{\kappa }{2}}a_\alpha \sqrt{\kappa }a_{i\alpha },\text{ }(\alpha =A,B),`$ with the following solution $$a_\alpha \left(t\right)=a_\alpha \left(0\right)e^{\frac{\kappa }{2}t}\sqrt{\kappa }_0^te^{\frac{\kappa }{2}\left(tt^{}\right)}a_{i\alpha }\left(t^{}\right)𝑑t^{}.$$ (4) When $`\kappa t`$ is considerably larger than $`1`$, from Eqs. (2) and (3), it follows that $`a_Aa_B`$ $`=`$ $`\sqrt{N\left(N+1\right)},`$ (4) $`a_\alpha ^{}a_\alpha `$ $`=`$ $`N,\text{ }\left(\alpha =A,B\right),`$ () $`a_\alpha a_\alpha ^{}`$ $`=`$ $`\left(N+1\right),\text{ }\left(\alpha =A,B\right).`$ (5) On the other hand, we know that two modes driven by a white noise are in Gaussian states at any time. A Gaussian state with the correlations (4) is necessarily a pure two-mode squeezed state. So the steady state of the cavity modes $`a_A`$ and $`a_B`$ is $$|\mathrm{\Psi }_{12}=S_{AB}\left(r\right)|\text{vac}_{AB},$$ (6) where the squeezing operator $`S_{AB}\left(r\right)=\mathrm{exp}\left[r\left(a_A^{}a_B^{}a_Aa_B\right)\right]`$ and the squeezing parameter $`r`$ is determined by $`\mathrm{coth}(r)=\sqrt{N+1}.`$ Next we include some important sources of noise in the state generation process. The noise includes the losses in the NOPA cavity and the light transmission loss from the NOPA cavity to the cavities A and B. With a small loss rate $`\eta _0\kappa _c`$ for the modes $`c_A`$ and $`c_B`$ in the NOPA cavity, the Langevin equation (1) is replaced by $`\underset{A}{\overset{.}{c}}`$ $`=`$ $`ϵc_B^{}{\displaystyle \frac{\kappa _c+\eta _0}{2}}c_A\sqrt{\kappa _c}c_{iA}\sqrt{\eta _0}v_{iA},`$ (6) $`\underset{B}{\overset{}{\stackrel{.}{c}}}`$ $`=`$ $`ϵ^{}c_A{\displaystyle \frac{\kappa _c+\eta _0}{2}}c_B^{}\sqrt{\kappa _c}c_{iB}^{}\sqrt{\eta _0}v_{iB}^{},`$ () where $`v_{iA}`$ and $`v_{iB}`$ are standard vacuum white noise, and the NOPA outputs are still given by $`c_{o\alpha }=c_{i\alpha }+\sqrt{\kappa _c}c_\alpha `$ $`\left(\alpha =A,B\right)`$. On the other hand, the transmission loss of light can be described by $$a_{i\alpha }=c_{o\alpha }\sqrt{e^{\eta _\alpha \tau }}+v_\alpha \sqrt{1e^{\eta _\alpha \tau }},\text{ }\left(\alpha =A,B\right),$$ (7) where $`\tau `$ is the transmission time, $`\eta _A`$ and $`\eta _B`$ are respectively the transmission loss rates for the outputs $`c_{oA}`$ and $`c_{oB}`$, and $`v_A`$ and $`v_B`$ are standard vacuum white noise. From Eqs. (6) and (7), it follows that the inputs for the cavities A and B have the following correlations $`a_{iA}\left(t\right)a_{iB}\left(t^{}\right)`$ $`=`$ $`\sqrt{N^{}\left(N^{}+1\right)}e^{\frac{\eta _A^{}+\eta _B^{}}{2}\tau }\delta \left(tt^{}\right),`$ $`a_{i\alpha }^{}\left(t\right)a_{i\alpha }\left(t^{}\right)`$ $`=`$ $`N^{}e^{\eta _\alpha ^{}\tau }\delta \left(tt^{}\right),\text{ }\left(\alpha =A,B\right),`$ $`a_{i\alpha }\left(t\right)a_{i\alpha }^{}\left(t^{}\right)`$ $`=`$ $`\left(N^{}e^{\eta _\alpha ^{}\tau }+1\right)\delta \left(tt^{}\right),\text{ }\left(\alpha =A,B\right).`$ where the total loss rates $`\eta _\alpha ^{}=\eta _\alpha +\frac{1}{\tau }\mathrm{ln}\left(1+\eta _0/\kappa _c\right)=\eta _\alpha +\eta _0/\left(\kappa _c\tau \right)`$ $`\left(\alpha =A,B\right)`$, and the parameter $`N^{}=\left|ϵ\right|^2\left(\kappa _c+\eta _0\right)^2/\left(\frac{\left(\kappa _c+\eta _0\right)^2}{4}\left|ϵ\right|^2\right)^2N`$. The steady state of the two cavity modes $`a_A`$ and $`a_B`$ is thus a Gaussian state with the non-zero correlations given by $`a_Aa_B`$ $`=`$ $`\sqrt{N\left(N+1\right)}e^{\frac{\eta _A^{}+\eta _B^{}}{2}\tau },`$ (8) $`a_\alpha ^{}a_\alpha `$ $`=`$ $`Ne^{\eta _\alpha ^{}\tau },\text{ }\left(\alpha =A,B\right),`$ () $`a_\alpha a_\alpha ^{}`$ $`=`$ $`\left(Ne^{\eta _\alpha ^{}\tau }+1\right),\text{ }\left(\alpha =A,B\right).`$ (9) The Gaussian state is completely determined by these correlations. The Gaussian state (8) can be equivalently described as the solution at time $`t=\tau `$ of the following master equation $`\stackrel{.}{\rho }`$ $`=`$ $`\eta _A^{}\left(a_A\rho a_A^{}{\displaystyle \frac{1}{2}}a_A^{}a_{A1}\rho {\displaystyle \frac{1}{2}}\rho a_A^{}a_A\right)`$ () $`+\eta _B^{}\left(a_B\rho a_B^{}{\displaystyle \frac{1}{2}}a_B^{}a_B\rho {\displaystyle \frac{1}{2}}\rho a_B^{}a_B\right)`$ with the initial state $`\rho \left(0\right)=|\mathrm{\Psi }_{AB}\mathrm{\Psi }|`$, where $`|\mathrm{\Psi }_{AB}`$ is defined by Eq. (5). This equivalence simplifies the physical picture in section 4, where we will use the master equation (9) to describe the state generation noise. ## III Entanglement concentration of pure two-mode squeezed states In the above, we have shown how to generate continuous partially entangled states between two distant cavities. In the case of no noise in the state generation process, the cavities are in a pure two-mode squeezed state. In this section, we will show how to concentrate continuous variable entanglement, that is, starting from several pairs of continuous entangled states, we want to get a state with more entanglement through only local operations. The section is divided into two parts. The first part describes the purification protocol for two entangled pairs, and the second part extends the protocol to include multiple pairs. ### A Concentration of two entangled pairs Assume now we have two cavities $`A_1,A_2`$ and $`B_1,B_2`$ on each side. Each pair of cavities $`A_i,B_i`$ $`\left(i=1,2\right)`$ are prepared in the state (5), which is now denoted by $`|\mathrm{\Psi }_{A_iB_i}`$. $`|\mathrm{\Psi }_{A_iB_i}`$, expressed in the number basis, has the form $$|\mathrm{\Psi }_{A_iB_i}=\sqrt{1\lambda ^2}\stackrel{\mathrm{}}{\underset{n=0}{}}\lambda ^n|n_{A_i}|n_{B_i},$$ (11) where $`\lambda =\mathrm{tanh}\left(r\right)`$. Equation (10) is just the Schmidt decomposition of the state $`|\mathrm{\Psi }_{A_iB_i}`$. For a pure state, the entanglement is uniquely quantified by the von Neumann entropy of the reduced density operator of its one-component. The entanglement of the state (10) is thus expressed as $$E\left(|\mathrm{\Psi }_{A_iB_i}\right)=\mathrm{cosh}^2r\mathrm{log}\left(\mathrm{cosh}^2r\right)\mathrm{sinh}^2r\mathrm{log}\left(\mathrm{sinh}^2r\right).$$ (12) The joint state of the two entangled pairs $`A_1,B_1`$ and $`A_2,B_2`$ is simply the product $`|\mathrm{\Psi }_{A_1B_1A_2B_2}`$ $`=`$ $`S_{A_1B_1}\left(r\right)|\text{vac}_{A_1B_1}S_{A_2B_2}\left(r\right)|\text{vac}_{A_2B_2}`$ (13) $`=`$ $`\left(1\lambda ^2\right){\displaystyle \stackrel{\mathrm{}}{\underset{j=0}{}}}\lambda ^j\sqrt{1+j}|j_{A_1A_2B_1B_2},`$ () where $`|j_{A_1A_2B_1B_2}`$ is defined as $$|j_{A_1A_2B_1B_2}=\frac{1}{\sqrt{1+j}}\stackrel{j}{\underset{n=0}{}}|n,jn_{A_1A_2}|n,jn_{B_1B_2}.$$ (14) We now perform a local QND measurement of the total photon number of the two cavities $`A_1,A_2`$. There have been several proposals for doing QND measurements of the photon number, and in section 5, we will describe a cavity scheme for realizing the QND measurement of the total photon number of two local cavities. Here we simply assume this type of measurement can be done. After the QND measurement of the total number $`n_{A_1}+n_{A_2}`$, the state $`|\mathrm{\Psi }_{A_1B_1A_2B_2}`$ is collapsed into $`|j_{A_1A_2B_1B_2}`$ with probability $$p_j=\left(1\lambda ^2\right)^2\lambda ^{2j}\left(j+1\right).$$ (15) The state $`|j_{A_1A_2B_1B_2}`$ is a maximally entangled state between the two parties $`A_1,A_2`$ and $`B_1,B_2`$ in a $`\left(j+1\right)\times \left(j+1\right)`$-dimensional Hilbert space, and its entanglement is $$E\left(|j_{A_1A_2B_1B_2}\right)=\mathrm{log}\left(j+1\right).$$ (16) If $`E\left(|j_{A_1A_2B_1B_2}\right)>E\left(|\mathrm{\Psi }_{A_iB_i}\right)`$, i.e. $`j>\frac{\left(\mathrm{cosh}\left(r\right)\right)^{\mathrm{cosh}\left(r\right)}}{\left(\mathrm{sinh}\left(r\right)\right)^{\mathrm{sinh}\left(r\right)}}1`$, we get a two-party state with more entanglement. The quantity $`\mathrm{\Gamma }_j=\frac{E\left(|j_{A_1A_2B_1B_2}\right)}{E\left(|\mathrm{\Psi }_{A_iB_i}\right)}`$ defines the entanglement increase ratio. Fig. 2 shows the probability of success versus entanglement increase ratio for some typical values of the squeezing parameter. An interesting feature of this entanglement purification protocol is that with any measurement outcome $`j0`$, we always get a useful maximally entangled state in some finite Hilbert space, though the entanglement of the outcome state $`|j_{A_1A_2B_1B_2}`$ does not necessarily exceed that of the original state $`|\mathrm{\Psi }_{A_iB_i}`$ if $`j`$ is small. The state $`|j_{A_1A_2B_1B_2}`$ involves two pairs of cavities. If one wants to transfer the entanglement to a single pair of cavity modes, one can make a phase measurement of the cavity mode $`A_2`$. There have been some proposals for doing a phase measurement . A phase measurement of the mode $`A_2`$ with the measurement outcome $`\varphi `$ will convert the state $`|j_{A_1A_2B_1B_2}`$ to the following maximally entangled state of a single pair of cavity modes $$|j_{A_1A_2}=\frac{1}{\sqrt{1+j}}\stackrel{j}{\underset{n=0}{}}e^{i\left(jn\right)\varphi }|n_{A_1}|n_{B_1}.$$ (17) ### B Concentration of multiple entangled pairs The above protocol can be extended straightforwardly to simultaneously concentrate entanglement of multiple cavity-pairs. Simultaneous concentration of multiple entangled pairs is much more effective that the entanglement concentration two by two. Assume that we have $`m`$ cavity-pairs $`A_1,B_1`$, $`A_2,B_2`$, $`\mathrm{}`$ and $`A_m,B_m`$. Each pair of cavities $`A_i,B_i`$ is prepared in the state (10). The joint state of the $`m`$ entangled pairs can be expressed as $`|\mathrm{\Psi }_{\left(A_iB_i\right\}}`$ $`=`$ $`|\mathrm{\Psi }_{A_1B_1}|\mathrm{\Psi }_{A_2B_2}\mathrm{}|\mathrm{\Psi }_{A_mB_m}`$ (18) $`=`$ $`\left(1\lambda ^2\right)^{\frac{m}{2}}{\displaystyle \stackrel{\mathrm{}}{\underset{j=0}{}}}\lambda ^j\sqrt{f_j^{\left(m\right)}}|j_{\left(A_iB_i\right\}},`$ () where $`\left(A_iB_i\right\}`$ is abbreviation of$`A_1,B_1`$, $`A_2,B_2,`$ $`\mathrm{},`$ $`A_m,B_m`$, and the normalized state $`|j_{\left(A_iB_i\right\}}`$ is defined as $$|j_{\left(A_iB_i\right\}}=\frac{1}{\sqrt{f_j^{\left(m\right)}}}\underset{\begin{array}{c}i_1,i_2,\mathrm{},i_m\\ i_1+i_2+\mathrm{}+i_m=j\end{array}}{}|i_1,i_2,\mathrm{},i_m_{\left(A_i\right\}}|i_1,i_2,\mathrm{},i_m_{\left(B_i\right\}}$$ (19) The function $`f_j^{\left(m\right)}`$ in Eqs. (17) and (18) is given by $$f_j^{\left(m\right)}=\frac{\left(j+m1\right)!}{j!\left(m1\right)!}=\left(\begin{array}{c}j+m1\\ m1\end{array}\right).$$ (20) To concentrate the entanglement, we perform a QND measurement of the total photon number $`n_{A_1}+n_{A_2}+\mathrm{}+n_{A_m}`$. This measurement projects the state $`|\mathrm{\Psi }_{\left(A_iB_i\right\}}`$ onto a two-party maximally entangled state $`|j_{\left(A_iB_i\right\}}`$ with probability $$p_j^{\left(m\right)}=\left(1\lambda ^2\right)^m\lambda ^{2j}f_j^{\left(m\right)}.$$ (21) The entanglement of the outcome state $`|j_{\left(A_iB_i\right\}}`$ is given by $$E\left(|j_{\left(A_iB_i\right\}}\right)=\mathrm{log}\left(f_j^{\left(m\right)}\right).$$ (22) Similarly, $`\mathrm{\Gamma }_j=E\left(|j_{\left(A_iB_i\right\}}\right)/E\left(|\mathrm{\Psi }_{A_iB_i}\right)`$ defines the entanglement increase ratio, and if $`\mathrm{\Gamma }_j>1`$, we get a more entangled state. For four pairs, the probability of success versus entanglement increase ratio is shown in Fig. 3. There appears a peak in the probability curve for some entanglement increase ratio between $`2`$ and $`3`$. To measure how efficient the scheme is, we define the entanglement transfer efficiency $`\mathrm{{\rm Y}}`$ with the expression $$\mathrm{{\rm Y}}=\frac{\stackrel{\mathrm{}}{\underset{j=0}{}}p_j^{\left(m\right)}E\left(|j_{\left(A_iB_i\right\}}\right)}{mE\left(|\mathrm{\Psi }_{A_iB_i}\right)}.$$ (23) It is the ratio of the average entanglement after concentration measurement to the initial total entanglement contained in the $`m`$ pairs. Obviously, $`\mathrm{{\rm Y}}1`$ should always hold. With the squeezing parameter $`r=0.5,`$ $`1.0`$ or $`1.5,`$ the entanglement transfer efficiency versus the number of pairs $`m`$ is shown in Fig. 4. From the figure, we see that the entanglement transfer efficiency is near to 1 for a large number of pairs. In fact, it can be proven that if $`m`$ goes to infinity, with unit probability we would get a maximally entangled state with entanglement $`mE\left(|\mathrm{\Psi }_{A_iB_i}\right).`$ To show this, we calculate the mean value and the variance of the distribution $`p_j^{\left(m\right)}`$, and find $`\overline{j}`$ $`=`$ $`{\displaystyle \frac{m\lambda ^2}{\left(1\lambda ^2\right)}},`$ (24) $`\overline{\left(\mathrm{\Delta }j\right)^2}`$ $`=`$ $`{\displaystyle \frac{m\lambda ^2}{\left(1\lambda ^2\right)^2}}.`$ () The results show that if $`m`$ tends to infinity, $`\sqrt{\overline{\left(\mathrm{\Delta }j\right)^2}}/\overline{j}0`$ and the distribution $`p_j^{\left(m\right)}`$ tends to a $`\delta `$-like function. Furthermore, around the mean value $`\overline{j}`$, the entanglement of the resulting state $`|\overline{j}_{\left(A_iB_i\right\}}`$ is $$E\left(|\overline{j}_{\left(A_iB_i\right\}}\right)\stackrel{m\mathrm{}}{}mE\left(|\mathrm{\Psi }_{A_iB_i}\right),$$ (25) so the entanglement transfer efficiency tends to unity. This proves that the purification method described above is optimal in the asymptotic limit ($`m\mathrm{}`$), analogous to the purification protocol presented in for the qubit case. For any finite number of entangled pairs, this purification protocol is more efficient than that in , since it takes advantage of the special relations between the coefficients in the two-mode squeezed state. ## IV Entanglement purification of mixed Gaussian continuous entangled states The assumption of noise-free preparation of partially continuous entangled states is not realistic. If we include the unavoidable light transmission loss and the NOPA cavity loss in the state generation process, in section 2 we have shown that we would get a mixed Gaussian continuous entangled state between two distant cavities. The state is described by the solution at the transmission time $`\tau `$ of the master equation (9), with the ideal two-mode squeezed state (10) at the beginning. If we want to establish $`m`$ entangled cavity-pairs $`A_1,B_1`$, $`A_2,B_2`$, $`\mathrm{}`$ and $`A_m,B_m`$, Eq. (9) can be extended directly to the following form $$\stackrel{.}{\rho }=i(H_{\text{eff}}\rho \rho H_{\text{eff}}^{})+\stackrel{m}{\underset{i=1}{}}(\eta _A^{}a_{A_i}\rho a_{A_i}^{}+\eta _B^{}a_{B_i}\rho a_{B_i}^{})$$ (26) where $`\rho `$ is the density operator of the whole $`m`$ entangled pairs with $`\rho \left(0\right)=|\mathrm{\Psi }_{\left(A_iB_i\right\}}\mathrm{\Psi }|`$, and the effective Hamiltonian $$H_{\text{eff}}=i\stackrel{m}{\underset{i=1}{}}\left(\frac{\eta _A^{}}{2}a_{A_i}^{}a_{A_i}+\frac{\eta _B^{}}{2}a_{B_i}^{}a_{B_i}\right).$$ (27) In Eqs. (25) and (26), we assumed that the total loss rates $`\eta _A^{}`$ and $`\eta _B^{}`$ are the same for the $`m`$ entangled pairs, but $`\eta _A^{}`$ and $`\eta _B^{}`$ may be different from each other. In this section, we will show how to distill entanglement from the kind of realistic continuous entangled states described by the solution of the master equation (25). There are two practical circumstances in which our entanglement purification protocol can be extended straightforwardly to generate maximally entangled states from the mixed Gaussian entangled states. We describe these two circumstances one by one. ### A Case of small state preparation noise Though the state preparation noise is unavoidable, in many cases it is reasonable to assume that it is quite small. We take $`\eta _A^{}\tau `$ and $`\eta _B^{}\tau `$ as small factors, and solve the master equation (25) perturbatively to the first order of these small factors. It is convenient to use the quantum trajectory language to explain the perturbative solution. In this language, to the first order of $`\eta _A^{}\tau `$ and $`\eta _B^{}\tau `$, the final normalized state of the $`m`$ entangled pairs is either (no jumps) $`|\mathrm{\Psi }^{\left(0\right)}_{\left(A_iB_i\right\}}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{p^{\left(0\right)}}}}e^{iH_{\text{eff}}\tau }|\mathrm{\Psi }_{\left(A_iB_i\right\}}`$ (28) $`=`$ $`{\displaystyle \frac{1}{\sqrt{p^{\left(0\right)}}}}\left(1\lambda ^2\right)^{\frac{m}{2}}{\displaystyle \stackrel{\mathrm{}}{\underset{j=0}{}}}\lambda ^je^{\frac{\eta _A^{}+\eta _B^{}}{2}\tau j}\sqrt{f_j^{\left(m\right)}}|j_{\left(A_iB_i\right\}},`$ () with probability $$p^{\left(0\right)}=\frac{\left(1\lambda ^2\right)^m}{\left(1\lambda ^2e^{\left(\eta _A^{}+\eta _B^{}\right)\tau }\right)^m}$$ (29) or (a jump occurred) $$|\mathrm{\Psi }^{\left(\alpha _i\right)}_{\left(A_iB_i\right\}}=\frac{1}{\sqrt{p^{\left(\alpha _i\right)}}}\sqrt{\eta _\alpha ^{}\tau }a_{\alpha _i}|\mathrm{\Psi }_{\left(A_iB_i\right\}},\text{ }\left(\alpha =A,B\text{ and }i=1,2,\mathrm{},m\right)$$ (30) with probability $`p^{\left(\alpha _i\right)}`$ $`=`$ $`\eta _\alpha ^{}\tau \text{ }_{\left(A_iB_i\right\}}\mathrm{\Psi }\left|a_{\alpha _i}^{}a_{\alpha _i}\right|\mathrm{\Psi }_{\left(A_iB_i\right\}}`$ (31) $`=`$ $`\overline{n}\eta _\alpha ^{}\tau ,`$ () where $`\overline{n}=_{\left(A_iB_i\right\}}\mathrm{\Psi }\left|a_{\alpha _i}^{}a_{\alpha _i}\right|\mathrm{\Psi }_{\left(A_iB_i\right\}}=\mathrm{sinh}^2\left(r\right)`$ is the mean photon number for a single mode. Similar to the pure state case, we also use QND measurements of the total photon number to distill entanglement from the mixed continuous state described by Eqs. (27)-(30). The difference is that now we perform QND measurements on both sides A and B. The measurement results are denoted by $`j_A`$ and $`j_B`$, respectively. We then compare $`j_A`$ and $`j_B`$ through classical communication, and keep the outcome state if and only if $`j_A=j_B`$. It is easy to show that the final state is a maximally entangled state in a finite dimensional Hilbert space. Let $`P_A^{\left(j\right)}`$ and $`P_B^{\left(j\right)}`$ denote the projections onto the eigenspace of the corresponding total number operator $`\stackrel{m}{\underset{i=1}{}}a_{A_i}^{}a_{A_i}`$ and $`\stackrel{m}{\underset{i=1}{}}a_{B_i}^{}a_{B_i}`$ with eigenvalue $`j`$, respectively. From Eqs. (27) and (29), it follows $`P_A^{\left(j\right)}P_B^{\left(j\right)}|\mathrm{\Psi }^{\left(0\right)}_{\left(A_iB_i\right\}}`$ $`=`$ $`|j_{\left(A_iB_i\right\}},`$ (32) $`P_A^{\left(j\right)}P_B^{\left(j\right)}|\mathrm{\Psi }^{\left(\alpha _i\right)}_{\left(A_iB_i\right\}}`$ $`=`$ $`0,\text{ }\left(\alpha =A,B\text{ and }i=1,2,\mathrm{},m\right)`$ () So if $`j_A=j_B`$, the outcome state is maximally entangled with entanglement $`\mathrm{log}\left(f_j^{\left(m\right)}\right)`$. The components (29) in the mixed density operator, which are not maximally entangled, are discarded through confirmation of the two-side measurement outcomes. Compared with the pure state case, the probability to get the entangled state $`|j_{\left(A_iB_i\right\}}`$ is now decreased to $$p_j^{}=\left(1\lambda ^2\right)^m\lambda ^{2j}f_j^{\left(m\right)}e^{\left(\eta _A^{}+\eta _B^{}\right)\tau j}.$$ (33) We also note that the projection operators $`P_A^{\left(j\right)}P_B^{\left(j\right)}`$ cannot eliminate the state obtained from the initial state $`|\mathrm{\Psi }_{\left(A_iB_i\right\}}`$ by a quantum jump on both sides A and B. The total probability for this kind of quantum jumps to occur is proportional to $`m^2\overline{n}^2\eta _A^{}\eta _B^{}\tau ^2`$. So the condition for small state preparation noise in fact requires $$m^2\overline{n}^2\left(\eta _A\tau +\eta _0/\kappa _c\right)\left(\eta _B\tau +\eta _0/\kappa _c\right)1.$$ (34) If the light transmission loss is the dominant noise, Eq. (33) reduces to $`m^2\overline{n}^2\eta _A\eta _B\tau ^21`$. ### B Case of asymmetric state preparation noise In the above purification protocol, we need classical communication (CC) to confirm that the measurement outcomes of the two sides are the same, and during this CC, we implicitly assume that the storage noise for the cavity modes is negligible. In fact, that the storage noise during CC is much smaller than the transmission noise is a common assumption made in all the entanglement purification schemes which need the help of repeated CCs . If we also make this assumption for continuous variable systems, there exists a simple purification protocol to generate maximally entangled states. We put the NOPA setup on the A side. After creation of ideal squeezed vacuum lights, we directly couple one output light of the NOPA to the cavity on side A without noisy propagation; and the other output of the NOPA is sent to the remote side B, through a long distance noisy transmission. This configuration of the setup is equivalent to setting the transmission loss rate $`\eta _A0`$ so that $`\eta _A^{}\eta _0/\left(\kappa _c\tau \right)`$. Note that the NOPA cavity loss rate $`\eta _0`$ is normally much smaller than the output coupling rate $`\kappa _c`$, so the total loss rate $`\eta _A^{}`$ can be much smaller than $`\eta _B^{}`$ in this case. The purification protocol now is exactly the same as that described in the previous case. We note that the component of the mixed density operator which is kept the projection $`P_A^{\left(j\right)}P_B^{\left(j\right)}`$ should subject to the same times of quantum jumps on each side A and B. We want this component is a maximally entangled state. This requires that the total probability for A and B to be subjected to the same nonzero number of quantum jumps should be very small. From Eq. (30), this total probability is always smaller than $`m\overline{n}\eta _A^{}\tau `$, no matter how large the transmission loss $`\eta _B\tau `$ is. So the working condition of the protocol in the asymmetric transmission noise case is $$m\overline{n}\eta _0/\kappa _c1.$$ (35) The transmission loss $`\eta _B\tau `$ can be above one. The probability of success for obtaining the maximally entangled state $`|j_{\left(A_iB_i\right\}}`$ is also given by Eq. (32). Before concluding this section, we remark that for continuous variable systems, the information carrier is normally light, and the assumption of storage with a very small loss rate is typically unrealistic. It is interesting to note that recently there have been proposals to store light in internal states of an atomic ensemble . If this turns out to be possible, the storage time for light can be greatly increased. Anyway, as was pointed out in , this purification method is in fact not essentially hampered by the difficulty to store light, since there is a simple posterior confirmation method to circumvent the storage problem. Note that the purpose to distill maximally entangled states is to directly apply them in some quantum communication protocol, such as in quantum cryptography or in quantum teleportation. So we can modify the above purification protocol by the following procedure: right after the cavity A attains its steady state, we make a QND measurement of the total excitation number on side A and get a measurement result $`j_A`$. Then we do not store the outcome state on side A, but immediately use it (e.g., perform the corresponding measurement as required by a quantum cryptography protocol). During this process, the modes $`B_i`$ are being sent to the distant side B, and when they arrive, we make another QND measurement of the total excitation number of the modes $`B_i`$ and get a outcome $`j_B`$. The resulting state on side B can be directly used (for quantum cryptography for instance) if $`j_A=j_B,`$ and discarded otherwise. By this method, we formally get maximally entangled states through posterior confirmation, and at the same time we need not store the modes on both sides. ## V QND measurements of the total photon number of several cavities The QND measurement of the total photon number plays a critical role in our entanglement purification protocol. There have been some proposals for making a QND measurement of the photon number in a single cavity , such as letting some atoms pass through the cavity, and measuring the internal or external degrees of freedom of the atoms . In this section, we propose a purely optical scheme for making a QND measurement of the total photon number contained in several cavities. The different optical modes interact with each other through cross phase modulation induced by a Kerr medium, and we use cavities to enhance this kind of interaction. As an illustrative example, in the following we will show how to measure the total photon number of two cavities. Extension of this scheme to include several cavities is straightforward. The schematic setup is depicted in Fig. 5. We want to make a QND measurement of the total photon number $`n_1+n_2`$ contained in the good cavities I and II, whose damping rate $`\kappa `$ is assumed to be very small. The cavities I and II, each with a Kerr type medium inside, are put respectively in a bigger ring cavity. The two ring cavities are assumed to damp at the same rate $`\gamma `$, and $`\gamma \kappa `$. A strong coherent light $`b_{i1}`$ is incident on the first ring cavity, whose output $`b_{o1}`$ is directed to the second ring cavity. The output $`b_{o2}`$ of the second ring cavity is continuously observed through homodyne detection, and we will show that under some realistic conditions, this detection gives a QND measurement of the total photon number operator $`n_1+n_2=a_1^{}a_1+a_2^{}a_2`$. The measurement model depicted in Fig. 5 is an example of a cascaded quantum system .The incident light $`b_{i1}`$ can be expressed as $$b_{i1}=b_{i1}^{}+g\sqrt{\gamma },$$ (36) where $`g\sqrt{\gamma }`$ is a constant driving field, and $`b_{i1}^{^{}}`$ is the standard vacuum white noise, satisfying $`b_{i1}^{}\left(t\right)b_{i1}^{}\left(t^{}\right)`$ $`=`$ $`0,`$ (37) $`b_{i1}^{}\left(t\right)b_{i1}^{}\left(t^{}\right)`$ $`=`$ $`\delta \left(tt^{}\right).`$ () The Hamiltonian for the Kerr medium is assumed to be $$H_i=\mathrm{}\chi n_ib_i^{}b_i,\text{ }\left(i=1,2\right),$$ (38) where $`b_1`$ and $`b_2`$ are the annihilation operators for the ring cavity modes, and $`\chi `$ is the cross-phase modulation coefficient. The self-phase modulation effects will be discussed in the next section and shown to be negligible under some realistic conditions. In the rotating frame, the Langevin equations describing the dynamics in the two ring cavities have the form $`\underset{1}{\overset{.}{b}}`$ $`=`$ $`i\chi n_1b_1{\displaystyle \frac{\gamma }{2}}b_1\sqrt{\gamma }b_{i1}^{}g\gamma ,`$ (39) $`\underset{2}{\overset{.}{b}}`$ $`=`$ $`i\chi n_2b_2{\displaystyle \frac{\gamma }{2}}b_2\sqrt{\gamma }b_{i2}`$ () The boundary conditions for the two ring cavities are described by $`b_{i2}`$ $`=`$ $`b_{o1}=b_{i1}^{}+g\sqrt{\gamma }+\sqrt{\gamma }b_1,`$ (40) $`b_{o2}`$ $`=`$ $`b_{i2}+\sqrt{\gamma }b_2.`$ () Assume $`\gamma \chi n_i,`$ $`\left(i=1,2\right)`$, and we take adiabatic elimination, i.e., let $`\underset{1}{\overset{.}{b}}=\underset{2}{\overset{.}{b}}=0`$ in Eq. (38), obtaining $`b_1`$ $``$ $`{\displaystyle \frac{2\left(g\gamma +\sqrt{\gamma }b_{i1}^{}\right)}{\gamma }}\left(1{\displaystyle \frac{2i\chi n_1}{\gamma }}\right),`$ (41) $`b_2`$ $``$ $`{\displaystyle \frac{2\left(g\gamma +\sqrt{\gamma }b_{i1}^{}\right)}{\gamma }}\left(1{\displaystyle \frac{4i\chi n_1}{\gamma }}{\displaystyle \frac{2i\chi n_2}{\gamma }}\right)`$ () Substituting the above result into Eq. (39), the final output field $`b_{o2}`$ is expressed as $$b_{o2}\frac{4ig\chi }{\sqrt{\gamma }}\left(n_1+n_2\right)+b_{i1}^{}+g\sqrt{\gamma }.$$ (42) Now we measure the $`X`$-component of the quadrature phase amplitudes of the output field $`b_{o2}`$ through a homodyne detection. The phase of the driving field $`g`$ is set according to $`g=i\left|g\right|`$. Suppose $`T`$ is the measuring time. What we really get is the integrated photon current over time $`T`$, which, divided by $`T,`$ corresponds to the following measuring operator $`X_T`$ $`=`$ $`{\displaystyle \frac{1}{T}}{\displaystyle _0^T}{\displaystyle \frac{1}{\sqrt{2}}}\left[b_{o2}\left(t\right)+b_{o2}^{}\left(t\right)\right]𝑑t`$ (43) $``$ $`{\displaystyle \frac{4\sqrt{2}\left|g\right|\chi }{\sqrt{\gamma }}}\left(n_1+n_2\right)+{\displaystyle \frac{1}{\sqrt{T}}}X_T^{\left(b\right)},`$ () where $`X_T^{\left(b\right)}=\frac{1}{\sqrt{2}}\left(b_T+b_T^{}\right)`$, and $`b_T`$, satisfying $`[b_T,b_T^{}]=1`$, is defined by $$b_T=\frac{1}{\sqrt{T}}_0^Tb_{i1}^{}\left(t\right)𝑑t.$$ (44) From Eq. (36), it follows that the defined mode $`b_T`$ is in a vacuum state. So the first term of the right hand side of Eq. (42) represents the signal which is proportional to $`n_1+n_2`$, and the second term represents the contribution of the vacuum noise. The distinguishability of this measurement is given by $$\delta n=\frac{\sqrt{\gamma }}{8\left|g\right|\chi \sqrt{T}}.$$ (45) If $`\delta n<1`$, i.e., if the measuring time $$T>\frac{\gamma }{64\left|g\right|^2\chi ^2},$$ (46) we perform an effective measurement of the total number operator $`n_1+n_2`$. During the measuring time $`T`$, the loss of the two cavities I and II should be negligible, which requires $$\kappa n_iT<1,\text{ }\left(i=1,2\right)$$ (47) Under this condition, $`n_1+n_2`$ is approximately a conserved observable, and we realize a QND measurement of the total photon number operator. The measurement projects the field in the cavities I and II to one of the eigenstates of $`n_1+n_2`$. Eqs. (45) and (46), combined together, determine the suitable choice for the measuring time. ## VI Influence of imperfections in the QND measurement We have shown how to perform a QND measurement of the total photon number. The scheme described above works under ideal conditions. For a real experiment, there are always many imperfections which should be considered. For example, the phase of the driving field may be unstable, and has a small variance; the damping rates and the cross phase modulation coefficients for different ring cavities may not be exactly the same; the Kerr media and the mirrors may absorb some light; self-phase modulation effects caused by the Kerr media may have some influence on the resulting state; there may be some loss of light from the first ring cavity to the second ring cavity; the efficiency of the detector is not unity. Of course, to realize a QND measurement of the total photon number, all the imperfections must be small. But the important question is how small these imperfections should be. In this section, we will deduce quantitative requirements for all the imperfections listed above. These calculations may be helpful for a future real experiment. We will consider these imperfections one by one. ### A Phase instability of the driving field Assume that the phase of the driving field $`g\sqrt{\gamma }`$ has a small variance $`\delta `$, i.e., $`g`$ is expressed as $`g=i\left|g\right|e^{i\delta }`$. Then, Eq. (42) is replaced by $$X_T\frac{4\sqrt{2}\left|g\right|\chi }{\sqrt{\gamma }}\left(n_1+n_2\right)+\frac{1}{\sqrt{T}}X_T^{\left(b\right)}\sqrt{2}\left|g\right|\delta \sqrt{\gamma },$$ (48) The last term of Eq. (47) represents the noise due to the phase instability of the driving field. It should be negligible compared with the signal, which requires $$\delta <\frac{4\chi }{\gamma }.$$ (49) On the other hand, we know that the squared phase variance $`\delta ^2`$ increases linearly with time, i.e. $`\delta ^2=\delta _tt`$, where $`\delta _t`$ is the increasing rate. The measuring time $`T`$ is bounded from below by Eq. (45), so the increasing rate of the phase instability of the driving field is required to satisfy $$\delta _t<\frac{1024\left|g\right|^2\chi ^4}{\gamma ^3}.$$ (50) Eq. (49) suggests it is easier to meet the requirement imposed by the phase instability with a strong driving field and a large cross phase modulation coefficient. ### B Imbalance between the ring cavities In the previous section, we assumed that the damping rates and the cross phase modulation coefficients are exactly the same for the two ring cavities. This may be impossible in a real experiment. Here we calculate the largest allowed imbalance between the two ring cavities. The damping rates and the cross phase modulation coefficients for the ring cavities are denoted by $`\gamma _1,`$ $`\gamma _2`$ and $`\chi _1,`$ $`\chi _2`$, respectively. The Langevin equations (38) and the boundary conditions (39) are replaced respectively by the following equations $`\underset{1}{\overset{.}{b}}`$ $`=`$ $`i\chi _1n_1b_1{\displaystyle \frac{\gamma _1}{2}}b_1\sqrt{\gamma _1}b_{i1}^{}g\gamma _1,`$ (51) $`\underset{2}{\overset{.}{b}}`$ $`=`$ $`i\chi _2n_2b_2{\displaystyle \frac{\gamma _2}{2}}b_2\sqrt{\gamma _2}b_{i2}`$ () $`b_{i2}`$ $`=`$ $`b_{o1}=b_{i1}^{}+g\sqrt{\gamma _1}+\sqrt{\gamma _1}b_1,`$ (52) $`b_{o2}`$ $`=`$ $`b_{i2}+\sqrt{\gamma _2}b_2.`$ () The final measured observable is expressed as $$X_T\frac{4\sqrt{2}\left|g\right|\chi _1}{\sqrt{\gamma _1}}\left(n_1+n_2\right)+\frac{1}{\sqrt{T}}X_T^{\left(b\right)}+4\sqrt{2}\left|g\right|\sqrt{\gamma _1}\left(\frac{\chi _2}{\gamma _2}\frac{\chi _1}{\gamma _1}\right)n_2,$$ (53) The last term of Eq. (52) represents the noise due to the unbalance between the ring cavities, which should be negligible compared with the signal, yielding $$\left|\frac{\chi _2\gamma _1}{\chi _1\gamma _2}1\right|<\frac{1}{n_2}.$$ (54) ### C Absorption and leakage of the light Light absorption by mirrors and Kerr media and light leakage through other mirrors of the ring cavities can be described by the same Langevin equation, which has the form $`\underset{1}{\overset{.}{b}}`$ $`=`$ $`i\chi n_1b_1{\displaystyle \frac{\gamma }{2}}b_1\sqrt{\gamma }b_{i1}^{}g\gamma {\displaystyle \frac{\beta _1}{2}}b_1\sqrt{\beta _1}c_{i1},`$ (55) $`\underset{2}{\overset{.}{b}}`$ $`=`$ $`i\chi n_2b_2{\displaystyle \frac{\gamma }{2}}b_2\sqrt{\gamma }b_{i2}{\displaystyle \frac{\beta _2}{2}}b_2\sqrt{\beta _2}c_{i2},`$ () where $`\beta _1`$ and $`\beta _2`$ are the light leakage (or absorption) rates of the first and second ring cavities, respectively, and $`c_{i1}`$ and $`c_{i2}`$ are the standard vacuum inputs. The boundary conditions for the ring cavities are still described by Eq. (39). The leaked (or absorbed) light fields $`c_{o1}`$ and $`c_{o2}`$ are expressed as $$c_{o\alpha }=c_{i\alpha }+\sqrt{\beta _\alpha }b_\alpha ,\text{ }\left(\alpha =1,2\right).$$ (56) The leakage (or absorption) of light may have two types of effects: First, it may destroy the balance between the two ring cavities; and second, the leaked light (55) may carry some information about $`n_1`$ (or $`n_2`$). Any information about $`n_1`$ (or $`n_2`$) will destroy the superposition of the different eigenstates of $`n_1`$ (or $`n_2`$), and thus lead to decoherence of the eigenstate of $`n_1+n_2`$ (Note that a eigenstate of $`n_1+n_2`$ is normally a superposition of the different eigenstates of $`n_1`$ (or $`n_2`$)). So we require that the information about $`n_1`$ (or $`n_2`$) carried by the leaked light should be completely masked by the vacuum noise. This is equivalent to require that the decoherence of the eigenstate of $`n_1+n_2`$ caused by the light leakage is negligible. To consider the first effect of the light leakage, we calculate the measured observable $`X_T`$, and find it has the form $$X_T\frac{4\sqrt{2}\left|g\right|\chi }{\sqrt{\gamma }}\left(n_1+n_2\right)+\frac{1}{\sqrt{T}}X_T^{\left(b\right)}+\frac{4\sqrt{2}\left|g\right|\chi }{\sqrt{\gamma }}\left(\frac{\beta _2^2}{\gamma ^2}\frac{\beta _1^2}{\gamma ^2}\right)n_2,$$ (57) The last term of Eq. (56) should be negligible compared with the signal, which requires $$\left|\beta _2^2\beta _1^2\right|<\frac{\gamma ^2}{n_2}.$$ (58) To consider the decoherence effect of the light leakage, we define a similar measuring operator $`X_T^{\left(\alpha \right)}`$ for the leaked light (55) $`X_T^{\left(\alpha \right)}`$ $`=`$ $`{\displaystyle \frac{1}{T}}{\displaystyle _0^T}{\displaystyle \frac{1}{\sqrt{2}}}\left[c_{o\alpha }\left(t\right)+c_{o\alpha }^{}\left(t\right)\right]𝑑t`$ (59) $``$ $`{\displaystyle \frac{8\sqrt{2}\left|g\right|\chi \sqrt{\beta _\alpha }\left(\alpha 1\right)}{\gamma }}\left(n_1+n_2\right)+{\displaystyle \frac{1}{\sqrt{T}}}X_T^{\left(c_\alpha \right)}`$ (60) $`{\displaystyle \frac{4\sqrt{2}\left|g\right|\chi \sqrt{\beta _\alpha }}{\gamma }}n_\alpha ,\text{ }\left(\alpha =1,2\right),`$ where $`X_T^{\left(c_\alpha \right)}`$, similar to $`X_T^{\left(b\right)}`$ defined below Eq. (42), are standard vacuum noise terms. The last term of Eq. (58) bears some information about $`n_\alpha `$, which should be completely masked by the vacuum noise term to make the decoherence effect negligible. This condition requires $$\frac{4\sqrt{2}\left|g\right|\chi n_\alpha }{\gamma }\sqrt{\beta _\alpha }<\frac{1}{\sqrt{2T}}.$$ (61) On the other hand, the measuring time $`T`$ is bounded from below by Eq. (45), which, combined with Eq. (59), yields the following requirement for the leakage rates $$\beta _\alpha <\frac{\gamma }{n_\alpha ^2},\text{ }\left(\alpha =1,2\right).$$ (61) Obviously, this is a much stronger requirement than that given by Eq. (57). We should mention that there is another kind of absorption by the Kerr medium, the absorption rate of which is proportional to the cavity photon number $`n_\alpha `$. This kind of absorption, usually termed two-photon absorption, cannot be described by Eq. (54). To incorporate the two-photon absorption, we add an imaginary part to the cross phase modulation coefficient $`\chi `$, i.e., $`\chi `$ is replaced by $`\chi +i\chi _i`$, where $`\chi _i`$ describes the two-photon absorption rate. The two-photon absorption should be negligible compared with the cross Kerr interaction, which requires $`\chi _i<\frac{\chi }{n_\alpha },`$ $`\left(\alpha =1,2\right).`$ ### D Self-phase modulation effects Normally, a Kerr medium also induces self-phase modulation effects. However, by a suitable choice of the resonance condition for the Kerr medium, the self-phase modulation effects can be made much smaller than the cross-phase modulation , then the self phase modulation interaction is basically negligible. Here, for completeness, we still calculate the influence of self-phase modulations. In fact, self-phase modulation of the ring cavity modes have no influence on the QND measurement. This modulation adds a term like $`i\chi _sb_i^{}b_ib_i`$ $`\left(i=1,2\right)`$ in the Langevin equation (38), where $`\chi _s`$ denotes the self-phase modulation coefficient for the ring cavity modes. We know that the ring cavity modes $`b_1`$ and $`b_2`$ are in steady states under adiabatic elimination, and to a good approximation $`b_i^{}b_i`$ can be replaced by $`b_i^{}b_i=4\left|g\right|^2`$. So the term $`i\chi _sb_i^{}b_ib_i`$ simply induces a constant phase shift for the output field $`b_{o2}`$, and it can be easily compensated by choosing the initial phase of the driving field $`g`$. Self phase modulation of the cavity modes $`a_1`$ and $`a_2`$ plays a more subtle role. First, it obviously has no influence on the QND measurement of $`n_1+n_2`$, but it influences the resulting state after the QND measurement. In the purification scheme for two entangled pairs (described in section III.A), if there is no self-phase modulation, the state after the QND measurement is given by Eq. (13); and if the self-phase modulation of the modes $`a_1`$ and $`a_2`$ is considered, the modulation Hamiltonian $`\mathrm{}\chi _s^{}n_i^2`$ $`\left(i=1,2\right)`$, in which $`\chi _s^{}`$ is the corresponding self phase modulation coefficient, will bring the resulting state into $$|j_{A_1A_2B_1B_2}^{}=\frac{1}{\sqrt{1+j}}\stackrel{j}{\underset{n=0}{}}e^{i\chi _s^{}t\left[n^2+\left(jn\right)^2\right]}|n,jn_{A_1A_2}|n,jn_{B_1B_2},$$ (62) where $`t`$ is the interaction time for the self phase modulation. It is important to note that the state (61) is still a maximally entangled state with entanglement $`\mathrm{log}\left(j+1\right)`$. In this sense, self-phase modulation effects have no influence on the entanglement purification, though the resulting state is changed. ### E Imperfect coupling from the first ring cavity to the second ring cavity If the coupling between the two ring cavities is not perfect, the relation $`b_{i2}=b_{o1}`$ is not valid any more, and should be replaced by $`b_{i2}`$ $`=`$ $`\sqrt{\mu }b_{o1}+\sqrt{1\mu }d_i,`$ (63) $`d_o`$ $`=`$ $`\sqrt{\mu }d_i+\sqrt{1\mu }b_{o1},`$ () where $`d_i`$ is the standard vacuum white noise, and $`d_o`$ represents the leaked light in the imperfect coupling. The quantity $`\mu `$ describes the coupling efficiency. This kind of imperfection is very similar to the light leakage (or absorption) described in subsection VI.C. The difference is that the imperfect coupling (62) does not cause any unbalance between the two ring cavities. The only restriction is that the decoherence effect induced by it should be negligible, which requires $$\mu >1\frac{1}{n_1^2}.$$ (64) Eq. (63) suggests that loss of light from the first to the second ring cavity should be very small. ### F Detector inefficiency The detector efficiency of course cannot attain $`1`$. For a detector with efficiency $`\nu `$, the real measured field $`b_{o2}^{}`$ has the following relation with the output of the second ring cavity $$b_{o2}^{}=\sqrt{\nu }b_{o2}+\sqrt{1\nu }e_i,$$ (65) where $`e_i`$ is the standard vacuum white noise. This imperfection is similar to the imperfect coupling considered in the previous subsection. But now the leaked light depends only on the operator sum $`n_1+n_2`$, and carries no information about the single cavity photon number $`n_1`$, so it does not induce any decoherence. The only role played by the detector inefficiency is that it decreases the signal by a factor $`\sqrt{\nu }`$, so Eq. (45) on the restriction of the measuring time is now replaced by $$T>\frac{\gamma }{64\nu \left|g\right|^2\chi ^2},$$ (66) Obviously, the detector inefficiency has no important influence on this QND measurement scheme. ## VII Summary and discussion In summary, we have given a detailed description of the purification protocol which generates maximally entangled states in a finite dimensional Hilbert space from two-mode squeezed states or from realistic Gaussian continuous entangled states. The nonlocal Gaussian continuous entangled states are generated by feeding two distant cavities with the outputs of the NOPA. The purification operation is based on a local QND measurement of the total photon number contained in several cavities. We have extensively analyzed a cavity scheme to do this QND measurement, and have deduced its working condition. Furthermore, we have discussed many imperfections existing in a real experiment, and deduced quantitative requirements for the relevant experimental parameters. In Tab. 1, we summarize the working conditions for the collective QND measurement, including the requirements for many types of imperfections. To realize the QND measurement, basically we need high finesse optical cavities and strong cross Kerr interaction media. A good example for the strong cross Kerr interaction is provided by the resonantly enhanced Kerr nonlinearity, which has been predicted theoretically and demonstrated in recent experiments . In those works, the Kerr medium is a low density cold trapped atomic gas, whose relevant energy level structure is represented by the four-state diagram shown in Fig. 6 with $`|1`$ being the ground state. The ring cavity mode $`b_i`$ with frequency $`\omega _b`$ is assumed to be resonant with the $`|1|3`$ transition, and the cavity mode $`a_i`$ with frequency $`\omega _a`$ ($`\omega _a`$ is quite different from $`\omega _b`$) is coupled to the $`|2|4`$ transition, but with a large detuning $`\mathrm{\Delta }_{42}`$. A nonperturbative classical coupling field with frequency $`\omega _c`$ resonant with the $`|2|3`$ transition creates an electromagnetically induced transparency (EIT) for the cavity fields $`a_i`$ and $`b_i`$. In this configuration, the one-photon absorption of the medium is eliminated due to quantum interference, and the cross Kerr nonlinearity is only limited by the two-photon absorption (the self Kerr nonlinearity is negligible provided that $`\left|\omega _a\omega _b\right|\mathrm{\Delta }_{42}`$). After adiabatically eliminating all the atomic levels, the cross phase modulation coefficient is given by $$\chi \frac{3\left|g_{13}\right|^2\left|g_{24}\right|^2}{\mathrm{\Omega }_c^2\mathrm{\Delta }_{42}}n_{\text{atom}},$$ (67) where $`g_{24}`$ and $`g_{13}`$ are the coupling coefficients between the atoms and the cavity modes $`a_i`$ and $`b_i`$, respectively, $`\mathrm{\Omega }_c`$ denotes the Rabi frequency of the coupling field, and $`n_{\text{atom}}`$ is the total atom number contained in the cavity. The two-photon absorption rate $`\chi _i`$ is connected with $`\chi `$ by the relation $`\chi _i/\chi =\gamma _{42}/\mathrm{\Delta }_{42}`$, where $`2\gamma _{42}`$ is the spontaneous emission rate from level $`|4`$ to level $`|2`$. To justify the adiabatic elimination, one requires that $`\frac{\left|g_{13}\right|^2n_{\text{atom}}}{\mathrm{\Omega }_c^2}<1`$ . As an estimation, if one takes $`\frac{\left|g_{13}\right|^2n_{\text{atom}}}{\mathrm{\Omega }_c^2}0.2,`$ $`g_{24}/2\pi 10`$MHz, $`\gamma _{42}/2\pi 30`$MHz, and $`\mathrm{\Delta }_{42}10\gamma _{42}`$, the coefficient $`\chi `$ is about $`\chi /2\pi 0.2`$MHz, and the two-photon absorption rate $`\chi _i0.1\chi `$. This value of the cross phase modulation coefficient $`\chi `$ is not large enough to realize a single-photon turnstile device , but it is enough for performing QND measurements of the photon number. For example, if the mean photon number $`n_1=n_2=\mathrm{sinh}^2\left(r\right)1.4`$ with the squeezing parameter $`r1.0`$, we choose the decay rates $`\kappa /2\pi 4`$MHz and $`\gamma /2\pi 100`$MHz (these values for decay rates are obtainable in current experiments ), and let $`g50`$ (for a cavity with cross area $`S0.5\times 10^4`$cm<sup>2</sup>, $`g50`$ corresponds to a coherent driving light with intensity about $`10`$mWcm<sup>-2</sup>). With the above parameters, all the requirements listed in Tab. 1 can be satisfied if we choose the measuring time $`T8`$ns. Note that the light speed can be much reduced in the EIT medium , so it is possible to get a reduced cavity decay rate $`\kappa `$ with the same finesse mirrors, and then more favorable parameters can be given for the QND measurement. Note also that a large Kerr nonlinearity based on EIT can also be obtained in other systems, such as trapping a single atom in a high finesse cavity . So the example discussed here is not the unique choice. We thank P. Grangier and S. Parkins for discussions. This work was supported by the Austrian Science Foundation, by the European TMR network Quantum Information, and by the Institute for Quantum Information. GG acknowledges support by the Friedrich-Naumann-Stiftung.
warning/0003/nlin0003045.html
ar5iv
text
# Using Topological Statistics to Detect Determinism in Time Series ## I Introduction This paper deals with the problem of determining whether the complex behaviour of a single time series may be explained in terms of a deterministic or a stochastic dynamical system. Although the idea was present from the very first days of the nonlinear time series analysis , methods explicitly aiming to detect determinism in time series have only been published rather recently. The fact that noise can mimic or mask deterministic (e.g. chaotic) behavior in classical measures of chaos (Lyapunov exponents, $`K_2`$ entropy and correlation dimension) has urged the need for more specific methods to discriminate deterministic from stochastic behaviour. Measures and vector fields , densities and trajectories , metric and topological , statistical and geometrical, these are, roughly speaking, different terms used to denominate the two broad approaches for investigating the chaotic behaviour of dynamical systems. It is a remarkable fact that most of the aforementioned methods to detect determinism are based in the study of trajectories or vector fields. Considering the noisy character of the inverse problem of time series analysis (from the point of view of nonlinear dynamics), the statistical approach must lie at the heart of any proposed methodology. This becomes relevant as soon as the system underlying a time series is ”excited” beyond the simple bifurcations, so that the geometrical information about the shape of the attractor, or the motion on it, is no longer available. This is also true in the case of stochastic dynamical system, where noise can make impossible to reconstruct the geometrical information . Moreover, trajectories and the geometry of the flow on the attractor are well defined in the case of a deterministic system, in contrast to the case of a stochastic dynamics. Bearing that in mind, we have developed a novel approach based in the study of the differentiability of the natural measure. The statistical character of our methodology is twofold. The first one concerns the use of the natural (or physical) measure. Differentiability of this invariant measure will be considered in the direction of the evolution, that is, along any typical trajectory. In this way, ”smoothness” of the trajectory and measure are considered in a single step. This is a general result of dissipative systems, regardless of its application to time series analysis. The second one is the use of statistics for analyzing mathematical properties , which allow us to test the differentiability of the measure in a statistical sense. In this way we can extend our methods to signal analysis. The procedure is based upon a formula that will be derived in the next section, and that will allow us to discriminate between deterministic and stochastic behaviours. This problem was tackled by us in whereas the basic idea of the methodology was proposed in . The paper is organized as follows. Section II is devoted to the mathematical background of our approach. At the end of the section we have included a brief summary of the concepts introduced in this section. This is addressed to the less inclined mathematical reader, which can skip most of the mathematical details given in the section. In Sec. II.A we introduce the fundamentals of our approach. The extension to stochastic dynamics is considered in Section II.B. General concepts about measure projections and time series are explained in section II.C, whereas Section II.D is devoted to a brief discussion of the statistics of topological properties used in this work. Section III is devoted to discuss the models to which we apply our method and the numerical procedures followed in each case. We also propose a variant of the surrogation method, partial surrogation, as a way to tune the degree of stochasticity. The results are presented in Section IV. The sensitivity of the time derivative of the measure to stochasticity, the dependence of the results on the embedding dimension and the application of our approach to mixed time series are illustrated on the models discussed in Section III. We also present a simple test on an experimental time series like the Belousov-Zhabotinski chemical reaction. ## II Theoretical Approach Here we discuss in detail the mathematical background of our method. The natural measure is first defined and the equation of motion that it obeys explicitly given. Then we show how stochasticity triggers a wild behavior of the time derivative of the natural measure. This is the feature upon which our proposal is based. In order to evaluate quantitatively this behavior we borrow the method proposed by Pecora et al . In particular we evaluate (statistically) the continuity of the logarithmic derivative of the natural measure. Those readers more interested in applications may directly go to the Summary at the end of the Section where the most important features of the method are highlighted with the aid of the equations discussed hereafter. ### A The Natural Invariant Measure along the trajectory Consider a dissipative dynamical system described by $`n`$-first-order differential equations $`\dot{𝐱}=𝐅(𝐱)`$. The corresponding flow $`f^t`$ maps a ”typical” initial condition $`𝐱_0`$ into $`𝐱(t)=f^t(𝐱_0)`$ at time $`t`$. Once transients are over, the motion settles over the attractor $`𝒜`$. Given such a system, we can define a probability space (measure space) $`(𝒜,,\mu )`$, where $``$ is the $`\sigma `$-algebra generated by the open sets of the invariant set $`𝒜`$, and $`\mu `$ is an invariant measure defined over the sub-sets of $``$ such that $`\mu :(0,1)`$. Of all the invariant measures which can be defined, only one is relevant from the point of view of the experimentalist or in computer simulations. This is the natural invariant measure, which gives the limiting distribution of almost all starting initial conditions. Using the indicator function $`1_B(𝐲)=1`$, if $`𝐲B_ϵ`$ and $`0`$ otherwise, we can define this measure for a set $`B_ϵ(𝐱)=\{𝐲:d(𝐱,𝐲)ϵ\}`$ as, $`\mu (B_ϵ(𝐱))=\underset{t\mathrm{}}{lim}{\displaystyle \frac{1}{t}}{\displaystyle _0^t}1_B(f^\tau (𝐲_0))𝑑\tau `$ for almost all $`𝐲_0`$ in the basin of attraction. The so–defined measure is invariant in the sense that it remains constant upon application of the evolution operator $`f^\tau `$, $`\mu (B_ϵ(𝐱))=\mu (f^\tau (B_ϵ(𝐱)).`$ In the present case what we calculate is $`\mu (B_ϵ(𝐱(t))`$ along the trajectory. This means that the evolution operator changes the point at which the set $`B_ϵ`$ is centered, namely, $`\mu (B_ϵ(f^\tau (𝐱)))`$, leaving unchanged the ball $`B_ϵ`$. This is the so–called Lagrangian evolution (see Fig. 1), that has to be distinguished from the Liouvillian evolution, $`\mu (B_ϵ(𝐱(t))f^\tau (\mu (B_ϵ(𝐱(t))`$, which, contracts the volume along the trajectory, as illustrated in Fig. 1. Calling $`\mu (𝐱(t))=\mu (B_ϵ(𝐱(𝐭)))`$, the ”material derivative” of the natural measure can be expressed as, $$\frac{d\mu (𝐱(t))}{dt}=\dot{𝐱}.\mu $$ (1) or $$\frac{d\mu (𝐱(t))}{dt}=.(\mu 𝐅)\mu .𝐅$$ (2) Eq. (2) only involves partial derivatives of $`\mu (𝐱(t))`$ along the trajectory, and since each trajectory in the attractor is contained in the support, $`\mu (𝐱(t))`$ is smooth along any trajectory, and is therefore well-defined. It is known that for axiom-A systems, hyperbolic and with a dense set of periodic orbits, it is possible to decompose the tangent bundle at each $`𝐱𝒜`$ in two linear sub-spaces, stable and unstable. It is also possible to define a measure in these systems, known as a SRB (Sinai-Ruelle-Bowen) which have the property of being smooth along the unstable manifold , due fundamentally to the stretching of the trajectories. On the other side, it is expected to have a wild behavior along the stable direction. In the present case, considering that Eq. (2) only involves derivatives along the trajectory and that in this direction no expansion or compression of the flux occurs (the associated Lyapunov exponent is equal to zero), we expect a smooth behaviour of the measure. ### B Stochastic Dynamics Consider now a stochastic dissipative dynamical system with additive noise, described by $`n`$-first-order differential equations, $`\dot{𝐱}=𝐅(𝐱)+\eta 𝐆(t)`$ (3) where $`\eta >0`$ is a small number (noise intensity) and $`𝐆(t)`$ is a vector of independently and identically distributed random Gaussian variables, of zero mean and correlations $`<G_i(t)G_j(t^{})>=\delta _{ij}\delta (tt^{})`$. A physical system will normally have a small level $`\eta `$ of random noise, so that it can be considered a stochastic process rather than a deterministic one. In a computer study, roundoff errors should play the role of the random noise. For suitable noise and $`\eta `$, the stochastic time evolution (3) has a unique stationary measure $`\mu `$ . This is the natural (or physical) invariant measure defined above. Inserting Eq. (3) in (1), $$\frac{d\mu (𝐱(t))}{dt}=(𝐅+\eta 𝐆(t)).\mu $$ (4) Whenever the vector field $`𝐅(𝐱)`$ can be expressed as $`𝐅(𝐱)=\varphi (𝐱)+𝐟(𝐱)`$, with $`𝐟(𝐱)`$ being orthogonal to the gradient term and having no divergence, the measure is given by : $$\mu (𝐱)=N\mathrm{exp}(\frac{\varphi (𝐱)}{\eta ^2})$$ (5) Then, introducing (5) in (4) we arrive at: $$\frac{d(\mathrm{ln}\mu (𝐱(t)))}{dt}=\frac{1}{\eta }(\frac{1}{\eta }|𝐅(𝐱)|^2+𝐆(t).𝐅(𝐱))$$ (6) In section II.A we discussed the smoothness of the measure along the trajectory for a deterministic system, at least in the case there exists a SRB measure. In the present case, that is, a stochastic process, it is possible to define a unique stationary measure that will tend to the SRB for $`\eta 0`$. This is the Kolmogorov Measure . At each time step of evolution, we add some noise (of amplitude $`\eta `$). If we repeat these operations (evolving by time evolution and putting some noise) again and again then we will get an invariant measure which is smooth along the unstable direction. This is because the deterministic part of the time evolution will improve the continuity of the density in the unstable direction by stretching, and roughening it in the other directions due to contraction. Thus, in the zero-noise limit $`\eta 0`$ we will get a measure $`\mu `$ that satisfies SRB conditions. Eq. (6) provide an alternative tool to investigate the findings of , according to which smoothness in phase space implies determinism in time series. For weak noise levels, the first term of the right hand side of (6) is dominant over the second term. In this case, smoothness in phase space implies ”continuity” in the left hand side of (6), or differentiability of the measure along the trajectory. On the other hand, in the case of strong noise levels, the second term is the dominant one and a wild behavior in the measure must be expected. This is so because the vector $`𝐆(t)`$ is uncorrelated with the actual position in the phase space. Although Eq. (6) was derived for a vector field obeying some restrictions (see above), our numerical results indicate that it can be applied to more general systems. ### C Time Series and Measure Projection In case that dynamical invariants are to be estimated from the knowledge of a single observable, we can rely upon Takens’ Delay Coordinate Map Theorem . According to it, the structure of the attractor (including differential information) can be preserved by using delay coordinate maps , as long as the embedding dimension $`m`$ is large enough to fully unfold the attractor structure. Furthermore, the Fractal Whitney Embedding Prevalence Theorem tells us that almost every smooth map will be an embedding (one-to-one and differential structure preserving) provided that $`m>2D`$, where $`D`$ is the box-counting dimension of the attractor. However, as shown in Refs. , when dealing with functions of the measure, embedding dimensions greater than the correlation dimension are enough. From the statistical point of view, we must ask how is the embedding process in the measure space $`(A,)`$ associated to the attractor $`A`$. That is, if $`\mu `$ is a measure in $`\mathrm{}^n`$ and $`\varphi :\mathrm{}^n\mathrm{}^m`$ is a function, we will have a projected or induced measure $`\mu _\pi =\varphi (\mu )`$ over a subset $`S\mathrm{}^m`$, where $`\varphi (\mu )(S)=\mu (\varphi ^1(S))`$. This is important because we are actually transforming the measure in the embedding process. Most of the existing literature is related with projections and reconstructions of the invariants sets and the dimensions of the corresponding probability measures . In our case, if $`\mu `$ is the physical measure describing the original system ($`\mu `$ is carried by an attractor in phase space) then the points in the reconstructed space are equidistributed with respect to the projected measure $`\mu _\pi `$ (except for particular cases avoided explicitly by the embedding theorems ). This is a sufficient condition which allows us to extract information about the original system working in its corresponding projection. If we have a time series, we can construct an $`m`$-dimensional vector: $`\overline{x_i}=(x_i,x_{i+\tau },\mathrm{},x_{i+(m1).\tau })`$ where $`\tau `$ is the time delay, chosen by one of the standard methods . Then, from a single set of observations, multivariate vectors in $`m`$-dimensional space are used to trace out the orbits of the system. Fig. 2 shows the concept of the reconstruction method. Explicitly shown is the reconstruction of the measure carried by the attractor $`\mu _\pi `$. ### D Statistics of Topological Properties Taken’s theorem , and its sequels, give a rigorous justification for state space reconstruction. The essence of the mathematical proof is that the trajectory formed from the time series is diffeomorphically related to the actual phase-space trajectory of the dynamical system. In order to test the mathematical properties embodied in the diffeomorfism, that is, continuity, differentiability, inverse differentiability and injectivity, Pecora et al. have developed a set of statistics intended to test quantitatively these properties. Their algorithms are of general use and can in particular be applied to test topological properties in any pair of sets of points. We will take advantage of this procedure by implementing the aforementioned algorithms to test numerically the continuity of the logarithmic derivative of the measure along the trajectory. Basically, the method is intended to evaluate, in terms of probability or confidence levels, whether two data sets are related by a mapping having the continuity property: A function $`f`$ is said to be continuous at a point $`𝐱_0`$ if $`ϵ>0,\delta >0`$ such that $`𝐱𝐱_0<\delta f(𝐱))f(𝐱_0)<ϵ`$. The results are tested against the null–hypothesis, specifically, the case in which no functional relation between points along the trajectory and the measure exists. This is done by means of the statistics proposed by Pecora et al $`\mathrm{\Theta }_{C^0}(ϵ)={\displaystyle \frac{1}{n_p}}{\displaystyle \underset{j=1}{\overset{n_p}{}}}\mathrm{\Theta }_{C^0}(ϵ,j)`$ (7) and $`\mathrm{\Theta }_{C^0}(ϵ,j)=1{\displaystyle \frac{p_j}{p_{max}}}`$ (8) where $`p_j`$ is the probability that all of the points in the $`\delta `$-set, around the point $`𝐱_j𝐱(t)`$, fall in the $`ϵ`$-set around $`\frac{d\mathrm{ln}\mu (𝐱_j)}{dt}`$. The likelihood that this will happen must be relative to the most likely event under the null hypothesis, $`p_{max}`$ (see reference ). When $`\mathrm{\Theta }_{C^0}(ϵ,j)1`$ we can confidently reject the null hypothesis, and assume that there exists a continuous function. As in the work of Pecora et al the $`ϵ`$ scale is relative to the standard deviation of the density time series, and thus, $`ϵ[0,1]`$. Plots of $`\mathrm{\Theta }_{C^0}(ϵ)`$ versus $`ϵ`$ can be used to quantify the degree of statistical continuity of a given function. In order to characterize the continuity statistics by means of a single parameter we have also calculated, $$\theta =_0^1\mathrm{\Theta }_{C^0}(ϵ)𝑑ϵ$$ (9) The limiting values of $`\theta `$, namely, 0 and 1, correspond to a strongly discontinuous and a fully continuous function, respectively. ### E Summary A naive test to quantify noise in signals is to check how smooth they are. As long as more noise contaminates the signal, more discontinuous it becomes. This is the case for example of additive noise, e.g. noise added to the signal. However, this is by no means a general rule. For instance, intrinsic noise, that is, noise added in the equation of motion, is not expected to affect the smoothness of the signal. This can be clearly seen in the case of surrogate time series: two time series (the original series and its surrogate) having the same correlation structure, may have very different underlying dynamics, i.e. one deterministic and the other stochastic. We can overcome the above drawback by using the trajectory of the system, instead of a single variable. Smoothness or continuity of the trajectory in phase space has been used before in this context. What we propose here is to use the distribution of points on the trajectory (or the natural measure) as a way to evaluate the the degree of noise in the system. Our proposal is based upon Eq. (6). This equation tells us that the amount of noise present in the system is directly related to the differentiability of the natural measure, evaluated along a typical trajectory. More noisy is the system, less differentiable it becomes. It remains the question of how to apply a fundamental analysis concept, like differentiability, to a magnitude defined numerically. The ”topological statistics” tools developed by Pecora et al came to our help. It is devised to evaluate in a statistical fashion topological properties of functions. Eq. (8) is the statistics used to test, against the null hypothesis, the degree of continuity of our function. As long as this statistic approaches unity we are confident that we have a continuous function. We must note that we have evaluated the continuity of the numerical derivative of the measure, which is equivalent to test differentiability, a computationally easier procedure. Continuity is directly related to the resolution with which we are looking at the function, that is $`ϵ`$, so the statistics is actually $`ϵ`$–dependent $`\mathrm{\Theta }_{C^0}(ϵ)`$. In order to provide us with an overall continuity test, we summed up over the whole range of $`ϵ`$ obtaining a single parameter $`\theta `$ (see Eq. (9)) which we use hereafter to characterize continuity. ## III Models and Numerical Procedures Here we discuss the various stochastic dynamical systems on which we have investigated the efficiency of our approach. The numerical procedures followed to reconstruct the space from time series related to some of the coordinates of those models and to evaluate the statistical differentiability of the natural measure are also described. We also discuss a variant of the surrogation process which consists of producing partially surrogate series. The method is a way to vary the degree of stochasticity of the time series. ### A The stochastic Van der Pol Oscillator The simplest case in which we can apply our ideas is in a nonlinear oscillator. We have used the Van der Pol (VdP) oscillator with an additive stochastic term. $`\dot{x}`$ $`=`$ $`y+\eta G_1`$ (10) $`\dot{y}`$ $`=`$ $`(x^21)yx+\eta G_2`$ (11) The parameter $`\eta `$ represents the noise level, and $`G_i(t)`$ are uncorrelated Gaussian noises, such that $`G_i(t)`$ Normal(0,$`\sigma `$), zero mean and standard deviation $`\sigma `$. Without loss of generality we take $`\eta =1`$, and tune the degree of noise only by the standard deviation $`\sigma `$. Numerical integration was carried out by means of an Euler algorithm. 20000 data points have been generated. Using the $`x`$-coordinate as our “experimental” time series, we have made a reconstruction with an embedding dimension in the range 2–20, and a $`\tau `$ lag of 10 (in sampling units). Every reconstruction has been rescaled to the unit hyper-square. Density time series have been obtained for each reconstruction. Starting with the first point in the reconstructed phase space, we follow the trajectory recording the density of points around each of the trajectory’s point. In order to estimate this density, we have used the Epanechnikov kernel , which, roughly speaking,“weighs” the points according to its distance to the center. This is preferable to the Gaussian kernel, which is of infinite support (and therefore has a lower computational efficiency) and the “square” kernel which gives equal status to the different points in the ball around the reference point. A parameter that has to be chosen carefully is the ball radius used to estimate the density. If the radius is too small, the low density regions will be practically depopulated and the measure will be underestimated. On the other hand, if the radius is too big, the estimation will capture points which are not really part of the neighborhood of the reference point. Of course, the ball radius is a function of the data points being used in the reconstruction process and must be chosen according to this fact. We have used a radius of the ball of 5% of the total attractor extent. In evaluating the continuity statistics, we average $`\mathrm{\Theta }_{C^0}(ϵ,j)`$ over $`n_p`$ points (see Eq. (6)) randomly distributed in the trajectory, typically $`10\%`$ of the total record. Fig. 3 shows a typical density time series from the $`x`$-coordinate of the Van der Pol oscillator. Albeit qualitative, the smooth behavior of this density along the trajectory is readily noted. ### B The stochastic Lorenz System The VdP oscillator will allow us to get a closer look at the procedure we are implementing. However, a nonlinear oscillator is a somewhat trivial example and we want to apply the method to more complicated cases. In fact, the ultimate objective of the methodology is to discriminate random behaviour from a deterministic one, and this is specially important in the case of chaotic behaviour. The Lorenz system with an additive stochastic term is an adequate choice. The related system of differential equations can be written as: $`\dot{x}`$ $`=`$ $`sx+sy+\eta G_1`$ (13) $`\dot{y}`$ $`=`$ $`y+rxxz+\eta G_2`$ (14) $`\dot{z}`$ $`=`$ $`bz+xy+\eta G_3`$ (15) The parameters used in the calculations are, $`s`$ = 10.0, $`r`$ = 28.0 y $`b`$ = 2.66, which give chaotic behaviour in the case $`\eta =0`$. $`G_i(t)`$ and $`\eta `$ were defined above. When not specified the results for the Lorenz system discussed in the following Section were obtained for an embedding dimension of 3, which is greater than the correlation dimension of the Lorenz system. Numerical integration of the Lorenz system was carried out by means of the Euler method. The time integration step was 0.01. Time series with 16384 data points and their respective surrogates were subsequently generated. The reconstruction was performed by the usual time–delay method , with a time delay given by the first zero of the autocorrelation estimate (10 in units of the integration step) on an embedded phase space of dimension 3. The natural measure $`\mu (𝐱(t))`$ along the trajectory was calculated by means of the Epanechnikov kernel density estimator with an sphere of radius 5% of the attractor extent. As in the VdP oscillator the continuity statistics was evaluated including up to 10% of the points in a given record. The sensitivity of the time derivative of the measure to stochasticity is illustrated in Fig. 4. This figure shows that whereas the surrogate of the $`x`$–coordinate time series is as ”smooth” as the original series, the time derivative of the logarithm of the measure is much more spiked in the surrogate than in the original series. ### C The Mackey–Glass model In order to investigate the effects of the embedding dimension we have also considered the high dimensional system introduced in . The dynamical system is described by means of the following delay–differential equation, $$\dot{x}=\frac{ax(t\delta }{1+x(t\delta )^c}bx(t)$$ (16) This equation has been proposed to model nonlinear feedback control in physiology. We use the set of parameters that gives an attractor dimension of $`7.5`$ , namely, $`a=0.2`$, $`b=0.1`$, $`c=10`$, and $`\delta =100`$. The resulting time series were analyzed with a time delay given by the first zero of the autocorrelation estimate, and the measure was evaluated on spheres of radius 10% of the attractor extent and the continuity statistics with 10% of the points in the total record. ### D Partial surrogation The surrogation process is a well established method in the context of nonlinear time series analysis . In typical applications a single time series is available. From this time series, an ensemble of the so-called surrogate series are generated that mimic certain properties of the original. For example, by simply scrambling the temporal order of the points in the original, one obtain surrogate time series which preserve the mean, variance, etc. One of the most popular methods of producing surrogate time series consists of shuffling the phases in the Fourier transform of the original data set . In this way, each value of the Fourier transform of the original data is multiplied by a random phase $`\mathrm{exp}(i\varphi )`$, with $`\varphi `$ random $`[0,2\pi ]`$ <sup>*</sup><sup>*</sup>*In order to get a real time series in the antitransformation we multiply symmetrically with respect to the center of the transform. The procedure generates a new time series with the autocorrelation structure of the original. Here we propose to introduce an additional factor, $`\mathrm{exp}(i\varphi \alpha )`$, with $`\alpha [0,1]`$. This factor allows us to control the degree of stochasticity by tuning the parameter $`\alpha `$. ## IV Results ### A Sensitivity of the time derivative of the measure to stochasticity In order to test the efficiency of the approach proposed here we have first evaluated the continuity statistics either on a coordinate or on the time derivative of the natural measure time series of the stochastic Van der Pol oscillator. Fig. 5 shows the continuity statistics for the $`x`$–coordinate and for the time derivative of the reconstructed natural measure with and without noise. There is almost no modification in the statistics of the coordinate upon noise addition. However, the statistics of the density time series reflects very clearly the presence of the stochastic term. This illustrates the efficiency and novelty of our approach and supports its application to more complex cases. We must remark that we are using the commonly named ”dynamical” noise, that is, a stochastic term added in the dynamical equations, instead of the ”measurement” noise, which is added after the ”clean” integration step. Preliminary results shows that an efficient method to discriminate both types of noise, can be achieved by using the continuity statistics over the density and the coordinates, but this deserve further research, and eventually will be published elsewhere. The results for the continuity statistics of the time derivative of the reconstructed measure from the $`x`$–coordinate of the Lorenz system are illustrated in Fig. 6. The results of Fig. 6a show that the time derivative of the measure in the original series is ”more continuous” (in a statistical sense) than its surrogate. Partial surrogation ($`10\%`$) decreases the degree of continuity of the time derivative of the measure in an extent lower than total surrogation, as expected. On the other hand, the results show that the continuity of the totally surrogated series shows almost no dependence on whether it has been derived from the original series or from a partially surrogated series (10 % surrogated). The results for the stochastic Lorenz system reported in Fig. 6b clearly show that the stochastic terms significantly decrease the statistical continuity of the time derivative of the measure. Surrogation of the stochastic series produces a further decrease of continuity, indicating that the series still has some degree of determinism. The degree of stochasticity of a time series can be quantified by calculating the integral of the continuity statistics as defined in Eq. (8). Fig. 7 a) shows how steeply $`\theta `$ decreases with the percentage of surrogation. Similarly, $`\theta `$ decreases with the standard deviation of the Gaussian noise in the stochastic Lorenz system (Fig. 7b), as expected. Thus, the magnitude $`\theta `$ can be used to evaluate the relative stochasticities of a set of experimental time series. ### B Dependence on the Embedding Dimension A point of crucial relevance is how the above results change with the embedding dimension $`m`$. We have investigated this question on the Lorenz system and on the high–dimensional system discussed in III.D . The results for the Lorenz system depicted in Fig. 8a show that $`\theta `$ decreases with the embedding dimension. This is a consequence of working with a fixed sphere radius for all $`m`$ and of the numerical noise that should increase with $`m`$. The decrease of $`\theta `$ is stronger in the surrogate series, although it is likely that the difference between the two should decrease for large enough $`m`$. In any case, the difference in $`\theta `$ between the original and the surrogate series changes only from 0.35 to 0.49 when $`m`$ is varied in the range 3–10. The behavior of $`\theta `$ in the high–dimensional system is far more intricate (see Fig. 8b). For $`m`$ well below the attractor dimension the measure for the surrogate series seems to be more continuous than that for the original series. The reason for this rather odd behavior has to be found in the heavy crossing of trajectories that occur at $`m`$ far below the attractor dimension . In those cases, surrogation seems to have a smoothing effect. Instead, for $`m>7`$ the behavior is similar to that of the Lorenz system, although the difference between the original and the surrogated records is substantially smaller. This point further study that is actually in progress. Fig. 9 illustrates the dependence of the continuity statistics of the measure on the embedding dimension $`m`$ for the stochastic Van der Pol oscillator. In this case the simplicity of the attractor gives an almost null dependence on $`m`$ for no noise or very low noise levels. When the noise is increased the behavior is in line with that found in the Lorenz system, namely, a decrease of $`\theta `$ as $`m`$ increases. This dependence on $`m`$ is more noticeable the greater the noise level. The study of the dependence of the continuity statistics of the measure derivative on the embedding dimension involves numerical difficulties that deserve some comments. In the case of low enough embedding dimension and moderate number of data points, the estimation can be achieved confidently for almost all the points in the trajectory. Of course, ”moderate” and ”long enough” are terms which depend on the attractor dimension. We think that the same criteria followed to get for example a reliable estimation of the correlation dimension , apply in this case. In our approach we have introduced noise explicitly, adding another factor to be considered in the estimation step. As it is well known, as ”extra” dimensions are included in the embedding process, noise populates them more or less uniformly. This is specially problematic in the case of ”simple” systems, like the VdP oscillator, as the trajectory in a stochastic oscillator may ”wander” far from the zero-noise limit cycle. In these excursions, the measure swept by the trajectory is unavoidably constant, because no other points are in the neiborhood of the evolution, except those which are time correlated. In such a case, a constant measure results, and thus, a high value of the continuity statistics. This effect is more noticeable as the embedding dimension is increased, because ”more” space is available (see Fig. 10). ### C Application to Mixed Time Series A distinctive feature of our method is the possibility of using it in different ranges of a given time series. In this way we can examine short records and evaluate its stochasticity. Bearing this in mind we have devised the following example: Suppose we have a time series which is half deterministic and half stochastic. Could our method discriminate both behaviors in the same time series?. In order to answer this question, we have generated a single time series (16384 points) with the first half coming from the $`x`$-coordinate of the deterministic Lorenz system, and the second half coming from its surrogate (100 % randomization) time series. We have applied the continuity statistics over four regions in the time derivative of the density record (two randomly selected in the first half and two in the second half). Fig. 11 shows the results. It is then clear that the statistics utilized here can discriminate stochastic from deterministic behaviour. Fig. 11 also shows the statistic for the whole time series (same number of reference points randomly selected along the time series). The results are midway between those for the stochastic and deterministic ranges. As another example of the use of our methodology we have investigated the effect of a noise burst in the Van der Pol oscillator. The dynamic equations are rewritten as, $`\dot{x}`$ $`=`$ $`y+A(t)\eta G_1`$ (17) $`\dot{y}`$ $`=`$ $`(x^21)yx+A(t)\eta G_2`$ (18) where $`A(t)=1`$ if $`t_1<t<t_2`$ and $`0`$ otherwise. The interval $`[t_1,t_2]`$ is a small interval where we ”turn on” the stochastic term. The idea behind this system is to test the capabilities of the method to detect the noise introduced. Using 20000 data points, we have used 500 consecutive points with the stochastic term added. In Fig. 12 we show a typical time series (and the measure time derivative) where noise has been turned on in the time interval 10000–10500, with a strength of $`\sigma =0.05`$. Fig. 13 shows Pecora et al. statistics applied to five regions of the whole series, one of them being the stochastic region. Again, our method clearly discriminates noisy and clean regions. ### D Application to Experimental Time Series In order to test our method in actual experimental time series, we have used the data set from the Belusov-Zhabotinski (BZ) chemical reaction . As shown by those authors the apparently random behaviour of the amplitude of the concentration of bromide ions can in fact be explained by deterministic laws. We will use here our approach to confirm the above finding. Using the bromide concentration time series, as in the work of Roux et al. we repeat the procedure explained above. We have used a $`\tau `$ = 30 (in sampling units) for the reconstruction and an embedding dimension of 4. Fig. 14 shows the continuity statistics for both the time derivative of the reconstructed measure using the BZ time series and for its 100 % surrogate. The large difference between the two and the rather high value of $`\mathrm{\Theta }_{C^0}`$ confirm our proposal in the sense that one can explain the behaviour of this record as the output of a deterministic dynamical system. ## V Concluding Remarks In brief, we have proposed a method to identify determinism in time series which exploits the continuity of the logarithmic time derivative of the natural measure along the trajectory, that is, its differentiability. The method is based upon a formula which explicitly shows the sensitivity of the measure to stochasticity. In the present work we have adapted the statistical method of Pecora et al. to investigate the continuity of the time derivative of the measure. Results of partially surrogated series and series derived from two stochastic dynamical systems and a high–dimensional system, clearly illustrate the suitability of the present method to the problem at hand. As we have shown, the method is very easily applied to any kind of time series, from the simplest one, as is the case of limit cycle oscillators, to the high–dimensional cases. Given a time series, and once done the standard reconstruction process, obtaining the density (or its time derivative) is a straightforward procedure. Then, one can use Pecora et al. approach over the whole time series or in selected pieces of it. A warning against a blind application of the method is in order, particularly in what concerns the evaluation of the continuity of the measure time derivative. A careful inspection of the density time series must be done before any further operation is performed. The dependence of the continuity statistics on the embedding dimension in low and high–dimensional systems, indicate that applications to real (experimental) time series would eventually require a thorough investigation of this point in each particular case, as is common in time series analysis. In any case, the fact that the method works reasonably well on short time series, supports its usefulness for the analysis of experimental series. We have shown this in a simple case as it is the Belousov-Zhabotinski reaction, confirming previous findings. In a broader sense, the application of continuity statistics over the density time series is a new aspect of the possibilities offered by the Lagrangian Measures . As we have shown previously, the use of more traditional tools on this density, such as Fourier Transforms or histograms, may help in extracting new information on the underlying dynamical system. ###### Acknowledgements. Thanks are due to E. Hernández, M. SanMiguel and R. Toral for many useful comments and suggestions. This work was supported by grants of the spanish CICYT (grant no. PB96–0085), the European TMR Network-Fractals c.n. FMRXCT980183, and of the Universidad Nacional de Quilmes. G. Ortega is member of CONICET Argentina and also thanks the ”Generalitat Valenciana” for support.
warning/0003/hep-ph0003176.html
ar5iv
text
# Does the quark-gluon plasma contain stable hadronic bubbles? ## I Introduction In a previous paper we calculated the free energy of a bubble of hadrons embedded in an extended quark-gluon plasma (QGP), and of a droplet of QGP embedded in an extended hadron phase, for parameters in the vicinity of the cosmological quark-hadron transition, i.e. the baryon chemical potential were set to zero. In Ref. we found, as Mardor and Svetitsky in , that the free energy of a hadron bubble of radius $`R`$ embedded in QGP possessed a minimum at radii of a few fm, even above the phase transition temperature. Thus, hadronization from QGP is strongly enhanced compared to the usual nucleation scenario, where an energy barrier has to be overcome before bubbles of hadrons can form and grow. An important ingredient in these calculations is the density of states of the relevant particles. In Ref. we advised a modification of the usual expressions arising from the multiple reflection expansion (MRE), and we saw that this modification (the MMRE) yielded more accurately the free energy than the MRE did when compared to a direct (but numerically heavier) sum-over-discrete-states calculation. Our modification of the multiple reflection expansion expressions is physically well motivated, in that it consists of a truncation of the density of states in such a way that we avoid the effects of a negative density of states, present in the usual MRE expressions. In this paper, we generalize the calculations to finite chemical potential, and we shall see that also in the case of non-zero chemical potential, the MMRE produces accurately the thermodynamic potential. We show that, as in the case of zero chemical potential, the thermodynamic potential of a hadron bubble embedded in QGP has a minimum at a radius of $`R1`$ fm, meaning that it is thermodynamically favorable for a QGP in equilibrium to spontaneously create bubbles of hadrons of about this size. However, the minimum in the thermodynamic potential found for non-zero chemical potential is less pronounced than in the zero chemical potential case considered in Ref. . The more general calculations presented in this paper should be relevant also to neutron star phenomenology as well as ultrarelativistic heavy ion collisions. In the following section we give an overview of the basic theory leading to the results of section III. Finally in Sec. IV, we summarize the conclusions. ## II Theoretical framework ### A The particle model #### 1 The quark-gluon plasma The model of the plasma phase is a close derivative of the MIT Bag model . It consists essentially of non-interacting quarks and gluons (also self-interaction of the gluons is neglected), bounded by MIT Bag-like boundary conditions. There are three quark flavors in our model: $`u`$ and $`d`$, which we consider massless, and $`s`$, to which we assign a mass of $`m_s=150`$ MeV. The other quarks are too heavy to be relevant in this analysis. We choose a critical temperature (the bulk phase equilibrium temperature at zero chemical potential) of $`T=150`$ MeV, which gives a Bag constant of $`B=(221`$ MeV$`)^4`$. The plasma may be inside a sphere, as in the usual MIT Bag (we shall refer to this configuration as a plasma droplet), or outside, i.e. in a “vacuum bubble” configuration. For a detailed review of this model, the reader may consult . We are mostly interested in the vacuum bubble situation, where the vacuum bubble is filled with a thermal mixture of hadrons as described later. For the density of states of the quarks and gluons we use a modification of the multiple reflection expansion (MRE) , advised in . This modification affects only the gluon density of states, which is then $$\rho _g(k,R)=\{\begin{array}{cc}0,\hfill & 0kR<0.832\hfill \\ \frac{V_{QGP}k^2}{2\pi ^2}\frac{4R}{3\pi },\hfill & kR0.832\hfill \end{array}$$ (1) for gluons inside a sphere (i.e. the MMRE($`R`$)), and $$\rho _g(k,R)=\{\begin{array}{cc}0,\hfill & 0kR<0.458\hfill \\ \frac{V_Hk^2}{2\pi ^2}+\frac{4R}{3\pi },\hfill & kR0.458\hfill \end{array}$$ (2) for the difference between the gluon density of states in a volume $`V_{\mathrm{}}`$ and in volume $`V_{\mathrm{}}V_H`$, (i.e. the MMRE($`R`$)). In we saw that this modified multiple reflection expansion (MMRE) description of the density of states made the free energy agree nicely with more direct calculations, i.e. summing over discrete energy levels. Before using the MMRE in the more general case under consideration in this paper, we performed similar checks of the MMRE against direct sum-over-discrete-states calculations. Also in this case, the MMRE proved to describe the density of states adequately and significantly better than the MRE, cf. Fig. 1. For a complete description of the MMRE density of states, we refer the reader to . #### 2 The hadron phase We include all hadrons with masses below $`1.2`$ GeV, taking only the volume part of the density of states into account. This is justified as the hadrons represent far fewer degrees of freedom than the QGP. Specifically, the density of states of a hadron species occupying a volume $`V_H=\frac{4\pi }{3}R^3`$ is $$\rho _H(k,R)=\frac{V_Hk^2}{2\pi ^2}.$$ (3) ### B Thermodynamics The thermodynamic potential is $$\mathrm{\Omega }(T,V,\{\mu _i\})=T\mathrm{ln}(Z(T,V,\{\mu _i\})),$$ (4) where $`Z(T,V,\{\mu _i\})`$ is the partition function $$Z(T,V,\{\mu _i\})=Tr\left\{e^{(_i\mu _i\mathrm{\Lambda }_i)/T}\right\}.$$ (5) Here, $``$ is the Hamiltonian, the $`\mathrm{\Lambda }_i`$’s are the symmetries of the Hamiltonian (the conserved quantities), and $`\{\mu _i\}`$ denotes the corresponding collection of chemical potentials. $`Tr\{..\}`$ means the sum over all diagonal elements of energy- and number-eigenstates. These are simultaneous eigenstates since $`[,\mathrm{\Lambda }_i]=0`$. The volume dependence enters via the energy levels, i.e. in the trace operation. In thermodynamic equilibrium, the configuration realized in Nature is the one which minimizes the thermodynamic potential, leading to the Gibbs conditions for phase equilibrium: $`T_1=T_2`$ (thermal equilibrium), $`\frac{\mathrm{\Omega }_1}{V_1}=\frac{\mathrm{\Omega }_2}{V_2}`$ (mechanical equilibrium), and $`\mu _i^{(1)}=\mu _i^{(2)}`$ (chemical equilibrium), where index $`1`$ and $`2`$ refer to the two phases. If there are no interactions between the particles, i.e. the energy of a particular state is the sum of the energies of the individual particles, then we can write the thermodynamic potential for a particular particle species enclosed in a spherical volume of radius $`R`$, as $$\mathrm{\Omega }(R,T,\mu )=gT_0^{\mathrm{}}𝑑k\rho (k,R)\mathrm{ln}(1\pm e^{(\sqrt{k^2+m^2}\mu )/T}),$$ (6) (upper sign for fermions, lower sign for bosons) where we further assume that the energy levels are sufficiently closely spaced that we may use a smoothed density of states, $`\rho (k,R)`$, (e.g. the MMRE) normalized such that $`_0^{\mathrm{}}𝑑k\rho (k,R)`$ is the total number of states in the volume $`V=\frac{4\pi }{3}R^3`$. The chemical potential entering in (6), is the combined chemical potential of that particle species $$\mu =\underset{i}{}\lambda _i\mu _i,$$ (7) $`\lambda _i`$ being the quantum expectation value of the symmetry operator $`\mathrm{\Lambda }_i`$ of the particle species in question (e.g. $`\lambda _{\mathrm{b}aryon}=1/3`$ for a quark). Finally, $`g`$ is an appropriate degeneracy factor. At the energies relevant in this analysis, the quarks and gluons can be considered point-like. This is not the case for the hadrons. We must take explicit account of the fact that the hadrons occupy a certain volume. A number of such excluded volume corrections have been discussed in the literature, but we shall not go into an analysis of the different suggestions. We choose the one proposed in , since it is very easy to implement and grasps much of the essential physics involved. According to , the “true” pressure of the hadron gas ($`p_H`$) is obtained from the ideal gas pressure ($`p_{H,id}`$) according to $$p_H=\frac{p_{H,id}}{1+ϵ_{H,id}/(4B)},$$ (8) where $$ϵ_{H,id}=\underset{i}{}ϵ_{i,id}$$ (9) is the total energy density of the ideal hadron gas, $$p_{H,id}=\underset{i}{}p_{i,id}$$ (10) is the total hadronic ideal gas pressure, ($`i\mathrm{h}adronspecies`$), and $`B`$ is the Bag constant. We note that some sort of excluded volume correction is essential, since in a point-like hadron gas, no transition to QGP will occur at low temperatures, even at arbitrarily high density. There is one more question about the hadrons we need to address before we proceed. This is the phenomenon of Bose-Einstein condensation, regarding the bosonic part of the hadron gas. A bosonic gas always has $`\mu _im_i`$, and when $`\mu _i=m_i`$ particles will start to “condense” into the lowest, zero-momentum state, meaning that there will be a macroscopic number of particles in this state. When the parameters are such that we would have $`\mu _im_i`$, the hadron species $`i`$ is removed from the above summations (9) and (10), since particles in a zero-momentum state do not contribute any pressure. ## III Results We shall consider both the case of equal chemical potentials, $`\mu _u=\mu _d=\mu _s\mu _q`$, and the case where the $`u`$ and $`d`$ quarks have a different chemical potential from that of the $`s`$ quark, $`\mu _u=\mu _d\mu _s`$. The first case is relevant when the time scales are such that the weak interactions maintain equilibrium between all three flavors of quarks (e.g. quark matter in neutron stars, neglecting the electron chemical potential as a first approximation), whereas the case of separate $`s`$ quark chemical potential is relevant to ultrarelativistic heavy ion collisions, the time scales here being such that the net number of $`s`$ quarks is conserved separately from that of the net number of light quarks. In our model, the two light quarks are both massless, and since we neglect electromagnetic interactions, there is no difference between $`u`$ and $`d`$ quarks. ### A Quark-hadron phase equilibrium To study the phase equilibrium between the hadrons and the QGP, we implement the Gibbs conditions on the volume (or bulk) part of the thermodynamic potential, i.e. also in the QGP density of states we only use the terms proportional to the volume. (We shall see later that the surface terms give rise to interesting features near bulk phase equilibrium.) In this way, we obtain the phase diagram in Fig. 2. The QGP phase (above the phase equilibrium lines) occupies a larger and larger region of the phase diagram as the $`s`$-quark chemical potential is raised. As one might have expected, the phase equilibrium line in the case of equal chemical potentials lies somewhere between the lines of varying $`\mu _s`$. ### B The thermodynamic potential In this section we look at the thermodynamic potential in the vicinity of equilibrium, now using the full expression of the density of states of the QGP (i.e. including the surface terms). In Figs. 36, there is a single chemical potential, $`\mu _q`$, common to all three quark flavors. First we show Fig. 3, the thermodynamic potential of a QGP droplet in equilibrium with an extended hadron phase near the bulk equilibrium point $`(\mu _q,T)=(320,106.1)`$MeV. The picture is as expected for a first order phase transition: Below the bulk equilibrium temperature no stable QGP droplet can form, and even somewhat above this temperature there is an energy barrier for the system to pass before the true minimum at $`R=\mathrm{}`$ can be reached. The surface terms are responsible for this energy barrier. Precisely at the phase transition temperature the volume terms cancel (by definition), and the thermodynamic potential goes to infinity as $`R^2`$ (the leading surface term). The more interesting conclusions are reached when one considers the reverse situation, namely a hadron bubble in equilibrium with an extended phase of QGP. Figs. 4, 5 and 6 show the thermodynamic potential of this configuration near different points on the phase equilibrium line. In all three cases the potential exhibits a minimum at a radius of $`R1`$ fm of approximately the same depth, even at temperatures above phase equilibrium temperature. Such a minimum would not be present in the standard textbook treatment of phase transitions, in which a first order phase transition usually is described in terms of a phenomenological free energy containing only volume- and (positive) surface terms. Thus, a QGP in thermodynamic equilibrium apparently contains bubbles of hadrons, and the transition from QGP to hadrons will proceed (in addition to ordinary bubble nucleation and subsequent expansion of these bubbles) by smooth expansion of the pre-formed bubbles. In addition to this smooth growth as the minimum tends to larger radii when the temperature drops, pre-formed bubbles can also grow discontinuously by passing the barrier from $`R>0`$ to some $`R`$ outside the barrier. This possibility resembles very much the ordinary nucleation scenario with a modified barrier height. At the parameters of Figs. 4, 5 and 6, the mean number of quarks in a hadron bubble of radius $`R=1`$ fm is typically 3–4, suggesting that formation of a real hadron is not a very rare event. However, the minimum being of modest depth, this interesting effect could be due to the model’s inaccurate representation of QCD. The fact that there is no energy barrier for a hadron bubble to form in QGP does not imply that the phase transition is second order, or a smooth cross-over. In the model adopted here, the transition is inherently first order, since there is an a priori difference in entropy between the two bulk phases, and thus a latent heat. As seen from e.g. Fig. 4 it is not energetically favorable for bulk hadronic matter to form at temperatures above the phase transition temperature $`T_0`$. But the formation of isolated bubbles of hadronic matter is favorable above $`T_0`$, so the transition will appear smoother than normally expected for a first order transition, with a more gradual release of latent heat. We now investigate the effect of letting the $`s`$ quark have its own chemical potential. To this end we show Figs. 7 and 8. In Fig. 7 the $`s`$ quark chemical potential is 4 times greater than the light quark chemical potential, whereas in Fig. 8 it is converse. The conclusion to be drawn from these figures is that a large $`s`$ quark chemical potential tends to wash out the minimum in the thermodynamic potential, present when the light quark chemical potential is less than or equal to the $`s`$ quark chemical potential. This behavior can be traced back to the positive surface tension contribution from the massive $`s`$-quarks to the thermodynamic potential. Finally, we underline that the interesting effect of bubble formation is due to the fact that the presence of a surface alters the density of states of particles. Although these finite-size effects are negligible at large radii of the bubble or droplet, they may have important implications for the phase transition as a whole. ## IV Conclusion We have seen that the thermodynamic potential of a hadron bubble embedded in quark-gluon plasma, exhibits a minimum at a radius $`R1`$ fm, even at temperatures somewhat above the bulk transition temperature. Thus, within the model described here, a homogeneous plasma of size larger than a few fermi is an impossibility. Regarding the relevance of these results in connection with current and forthcoming ultrarelativistic heavy ion collisions, we conclude that if indeed a quark-gluon plasma is formed in the course of such a collision, then (according to the model considered here) this plasma phase will contain bubbles of hadrons of radius $`R1`$ fm. If such hadronic bubbles form inside the plasma, the observable effects are likely to include some blurring of the plasma signatures, since this reduces the effective plasma volume. We emphasize that these conclusions may well be model dependent, inasmuch the minimum of the thermodynamic potential is of modest depth. However, the phenomenon of spontaneous creation of stable hadronic bubbles in a quark-gluon plasma, does seem to be well established within the the model discussed here . The main question is now, whether these hadronic bubbles are of physical nature, or merely an artifact of the model. Certainly, it would improve confidence in these results if other models of QCD, and eventually lattice calculations, were to yield similar results. ###### Acknowledgements. JM was supported in part by the Theoretical Astrophysics Center under the Danish National Research Foundation. We thank Michael Christiansen for useful discussions.
warning/0003/astro-ph0003461.html
ar5iv
text
# The quasar fraction in low–frequency selected complete samples and implications for unified schemes ## 1 Introduction Orientation-based unified schemes for radio-loud quasars and powerful radio galaxies play a key role in our understanding of these objects. Ever since the conception of the idea that powerful radio galaxies are simply quasars with their jet axes oriented away from our line of sight, such that the nuclear continuum and broad-line regions are obscured by a dusty torus or warped disc (Scheuer 1987; Barthel 1989), evidence has been sought to prove that these unification schemes are correct. Reviews of unification schemes for active galactic nuclei have been presented by Antonucci (1993) and Urry & Padovani (1995). Although the majority of observations are consistent with them, some observations have been used to cast doubt on their viability over all ranges of radio luminosities and redshifts. The narrow line emission in radio sources is observed to be emitted largely from beyond the obscuring material (e.g. McCarthy, Spinrad & van Breugel 1995; Hes, Barthel & Fosbury 1996) and therefore is independent of the jet axis orientation. Hence in this model one would expect radio galaxies and quasars to have similar narrow line luminosities. The strong positive correlation between the extended radio luminosities and narrow line luminosities of radio sources (Baum & Heckman 1989; Rawlings et al. 1989; Willott et al. 1999) means that the samples of radio galaxies and quasars to be compared must be matched in extended radio luminosity. At low redshift ($`z<0.8`$), there have been claims that quasars are observed to have \[OIII\] line luminosities a factor of 5-10 greater than radio galaxies of similar radio luminosities (Baum & Heckman 1989; Jackson & Browne 1990; Lawrence 1991), although the \[OII\] luminosities of radio galaxies and quasars at these redshifts are indistinguishable (Browne & Jackson 1992; Hes et al. 1996). These differences between \[OII\] and \[OIII\] have been interpreted as being due to partial obscuration of \[OIII\] since its higher ionization potential means it is likely to be emitted from closer to the nucleus than \[OII\]. However, Jackson & Rawlings (1997) have investigated the \[OIII\] luminosities of $`z>1`$ radio galaxies and quasars and find their distributions indistinguishable. Using a combined 7C/3CRR dataset Willott et al. (1999) find that quasars have more luminous narrow lines than radio galaxies at intermediate radio luminosities \[$`26<\mathrm{log}_{10}(L_{151}`$ / W Hz<sup>-1</sup>sr$`{}_{}{}^{1})<27`$\], but they are similar at higher radio luminosities; this result leads to different slopes for the narrow-line versus radio luminosity correlation for quasars and radio galaxies. Barthel (1989) showed that in the redshift range $`0.5<z<1.0`$ the linear size distributions and fraction of quasars in the 3CRR sample are consistent with all radio galaxies having their jet axes at an angle $`\theta _{\mathrm{trans}}>44^{}`$ from our line-of-sight and all the quasars having $`\theta _{\mathrm{trans}}<44^{}`$, where $`\theta _{\mathrm{trans}}`$ is the half-opening angle of the obscuring torus; a value of $`\theta _{\mathrm{trans}}=45^{}`$ has been adopted by many on the basis of this paper. However, at higher redshifts in the 3CRR sample, the fraction of quasars increases giving $`\theta _{\mathrm{trans}}60^{}`$ (e.g. Singal 1993). Note that because of the tight luminosity–redshift correlation inherent in a single flux-limited sample, this apparent correlation with redshift may be due instead to a correlation with radio luminosity. Singal (1996) used several large samples with differing radio flux-density limits to find that the quasar fraction in radio samples declines with decreasing radio flux-densities. However, due to the correlation between the optical and radio luminosities of steep-spectrum quasars (Serjeant et al. 1998; Willott et al. 1998a), radio-fainter samples will contain optically-fainter quasars. Therefore the quasar fraction in faint samples might be underestimated if only quasars brighter than a certain optical magnitude limit are identified, which may well be the case in some of these fainter samples used by Singal. Using the new 7C Redshift Survey, a low-frequency radio sample selected at a flux-density limit $`25\times `$ lower than the 3CRR sample with $`90\%`$ spectroscopic redshift completeness, we have previously shown that the quasar fraction does not depend strongly upon radio luminosity or redshift for $`1<z<3`$ (Willott et al. 1998b). Note that the low-frequency selection of the samples is crucial, because this ensures selection on extended radio flux which should be entirely isotropic. In this paper we use a larger sample to investigate how the quasar fraction of radio sources depends upon redshift, radio luminosity and narrow emission line luminosity and we consider the implications for orientation-based unified schemes. In a companion paper (Blundell, Rawlings & Willott in prep.) we will investigate the constraints placed on unified schemes by the radio properties of quasars and radio galaxies from the combined 7C/6C/3CRR dataset. We assume throughout this paper that $`H_{}=50\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$ and $`q_{}=0.5`$. ## 2 The sample Our sample consists of three completely–identified samples selected at (similar) low radio frequencies. These samples have different flux limits to provide broad coverage of the radio luminosity–redshift ($`L_{151}z`$) plane (see Figure 1). This enables the separation of evolutionary– and luminosity–dependent effects. Our faintest sample is the 7C Redshift Survey which is briefly described here (see also Willott et al. 1998a; Willott et al. in prep.; Blundell et al. in prep; Lacy et al. 1999). The 7C Redshift Survey includes all sources with flux-densities at 151 MHz $`S_{151}0.5`$ Jy in three selected regions of sky (7C-I, 7C-II and 7C-III with a total sky area of 0.022 sr). The sample contains 130 radio sources which have all been identified with an optical/near-IR counterpart. The spectroscopic completeness is $`90`$% with most of the remaining sources having redshifts well-constrained by optical/near-IR photometry (Willott, Rawlings & Blundell, in prep.). The bright sample used is the 3CRR sample of Laing, Riley & Longair (1983), which has complete redshift information for all 173 sources, selected with $`S_{178}10.9`$ Jy. 3C 231 (M82) is excluded because it is a nearby starburst galaxy and not a radio-loud AGN. 3C 345 and 3C 454.3 are flat-spectrum quasars which are excluded on the grounds of Doppler-boosting of compact components being responsible for raising their fluxes above the selection limit. One quasar from the 7C sample (5C7.230) was excluded for the same reason. The intermediate flux-density sample used is a revision of the 6C sample of Eales (1985) (see Rawlings, Eales & Lacy in prep., hereafter REL, for details). The flux limits of this sample are $`2.0S_{151}<3.93`$ Jy and the sky area covered is 0.103 sr. Only 2 of the 58 sources in this sample do not have redshifts determined from spectroscopy. One object is faint in the near–infrared ($`K>19`$), so we take it as a galaxy at $`z=2.0`$ in this paper. The other is relatively bright in the near-IR but very faint in the optical. Its red colour suggests a galaxy with a redshift in the range $`0.8<z<2`$ and we assume $`z=1.4`$ here. Full details of the sample, including optical spectra, will appear in REL. All the sources in two of the three 7C regions, 7C-I and 7C-II, and in the 6C sample have been imaged in the near–infrared $`K`$-band. This has enabled the discovery of a few lightly reddened quasars, which would otherwise have been identified as galaxies on the basis of optical spectroscopy (see e.g. Willott et al. 1998a). In addition we have obtained near–infrared spectroscopy of some possible reddened quasars in these samples to search for broad emission lines (Willott et al. in prep.). On the basis of the available imaging and spectroscopy, all sources have been classified into one of three categories: quasar, broad line radio galaxy (BLRG) or radio galaxy. Given the relatively small number of reddened quasars found in the 6C, 7C-I and 7C-II samples, we do not believe the lack of near–infrared data for the 7C-III sample has caused a statistically–significant number of quasars to be mis–classified. Spectra of the quasars and BLRGs in 7C-I and 7C-II are presented with a detailed discussion of the classification scheme in Willott et al. (1998a). Our total sample comprises 356 sources (one object, 3C 200, is common to both the 3CRR and 7C-II samples). We do not exclude compact symmetric objects (CSOs) from our analysis as previous authors have (e.g. Barthel 1989, Singal 1996), since we have no a priori reason to expect them not to be part of the unified schemes. In Section 3 we repeat our analysis excluding the CSOs to check this assumption. Sources with FRI radio structure are however excluded. The main reason for this was a practical one: most (24 out of 33) of the FRIs are from the 3CRR sample, and the existing spectrophotometry of these objects is inadequate for the quantitative investigations of this paper. However, as discussed fully in Section 4.5, the lack of quasars with FRI radio structure (e.g. Falcke, Gopal-Krishna & Biermann 1995), and the concentration of these objects at low $`L_{151}`$ and low $`z`$ (see Fig. 1) makes it easy to investigate the effects of the exclusion of FRIs on our study of the quasar fraction. Although the radio structures of the quasars 5C6.264 and 3C 48 are arguably FRI, we count them in this paper as FRIIs because of their high radio luminosities. Indeed all objects that could not unambiguously characterised as FRI or FRII (e.g. the ‘DA’ sources in the notation of Blundell et al. in prep.) are also counted as FRIIs in this study. Thus our total combined sample contains 323 objects deemed to be FRIIs. ## 3 The quasar fraction We now investigate the fraction of objects which show broad emission lines (referred to hereafter simply as the quasar fraction) in the complete radio samples. Broad line radio galaxies (broad line objects with $`M_B>23`$) are counted as quasars in this analysis, because they are most likely to be weak quasars viewed within the torus opening angle (Laing et al. 1994; Hardcastle et al. 1998; Fig. 6 of Willott et al. 1998a). However, objects with scattered broad lines seen only in polarised light (4 cases) are counted as galaxies, because the fact that the broad lines are not observed directly indicates that the nuclear regions are heavily obscured, most likely by the torus. Since spectropolarimetry is unavailable for most sources, there is clearly some danger that the split between weak and scattered-light quasars is imperfect. A further concern is the question of the classification of lightly-reddened broad-line objects in which it remains uncertain whether (a) the reddening is related to the torus or is caused by dust further from the nucleus, for example in the host galaxy of the quasar; and (b) whether, particularly given the inhomogeneous spectrophotometric dataset for the 3CRR sample, all such objects have yet to be discovered. Taking $`0\stackrel{<}{}A_V\stackrel{<}{}5`$ as a working definition of ‘light’ reddening, we have chosen to count all such objects known to us as quasars. We will return to this important point in Section 4.2. In Fig. 1 (left), we plot low-frequency radio luminosity $`L_{151}`$ against redshift $`z`$ for the combined sample of FRII sources. The $`L_{151}z`$ plane has been binned and the percentage of quasars in each bin shown, along with the associated Poisson errors. The first thing to note is that there are very few quasars at $`\mathrm{log}_{10}(L_{151}/`$ W Hz<sup>-1</sup> sr$`{}_{}{}^{1})<26.5`$. Excluding this region, the quasar fraction appears to increase slightly as a function of both redshift and radio luminosity, ranging from 0.3 to 0.5. However, with Poisson errors of $`\sigma 0.08`$ these differences can at best be called marginally–significant, and we will not consider them further. Note also that for the high-luminosity bins, the median luminosity in each bin changes with redshift, so the effects of redshift and luminosity have not been completely separated. Laing et al. (1994) proposed that by excluding low–excitation radio galaxies (LEGs; Hine & Longair 1979) which have very weak or absent emission lines and low–ionization narrow lines, the quasar fraction in 3CRR was not a function of luminosity (or redshift). Their classification for LEGs was that they have \[OIII\] to H$`\alpha `$ ratios of $`<0.2`$ and \[OIII\] equivalent widths of $`<3`$ Å. Since the radio galaxies in the combined samples here span a large range of redshift, Balmer and/or \[OIII\] lines are often not observed in optical spectra and a classification scheme such as this cannot be applied. Instead we separate sources into those with high and low \[OII\] emission line luminosities. The division between these classes we adopt is $`\mathrm{log}_{10}(L_{[\mathrm{OII}]}/\mathrm{W})=35.1`$. Below this line luminosity there are very few broad line objects in the complete samples (Willott et al. 1999). This cut in narrow emission line luminosity was chosen to be equivalent to a rest-frame \[OII\] equivalent width of 10 Å for a quasar with $`M_B=23`$<sup>2</sup><sup>2</sup>2The mean rest-frame equivalent width of the \[OIII\] line for BQS quasars in the study of Miller et al. (1992) is 30 Å. Assuming a ratio of 3:1 for the \[OIII\]:\[OII\] equivalent widths, this translates to 10 Å for \[OII\]. This value was used in Willott et al. (1999) to determine the strength of the photoionizing continuum in radio sources. However, the typical ratio between \[OII\] and \[OIII\] may actually be smaller than this, causing Willott et al to have overestimated the mean \[OII\] equivalent width by a factor of a few. Baker et al. (1999) finds a median \[OII\] equivalent width of 4 Å in the Molongo Quasar Sample. Also, in the LBQS composite quasar spectrum of Francis et al. (1991) the \[OII\] equivalent width is 2 Å, which would give a line luminosity of $`\mathrm{log}_{10}L_{[\mathrm{OII}]}=35.1`$ for a $`M_B=25`$ quasar.. Note that this division puts more sources in the low-luminosity category than those typically classified as LEGs. On the right-hand side of Fig. 1 we plot only sources with \[OII\] luminosities $`\mathrm{log}_{10}(L_{[\mathrm{OII}]}/\mathrm{W})35.1`$. The differences between the quasar fractions in the bins is reduced somewhat here. Now we find that (with the exception of the poorly–populated low radio luminosity bin) the quasar fraction is 0.40 in all the bins at the 1$`\sigma `$ level. The largest change has occurred in the low-redshift, intermediate luminosity bin which contains many weak-lined 3CRR galaxies. Note that the $`1z<2`$, intermediate luminosity bin has the lowest quasar fraction. A possible reason for this is that this bin contains most of the galaxies without lines in their optical spectra which may not all truly lie within this redshift range. Considering all 216 high emission line luminosity sources plotted on the right-hand side of Fig. 1, we find a quasar fraction of $`0.40\pm 0.03`$ implying a mean torus half-opening angle of $`53^{}\pm 3^{}`$ for the luminous population. There is likely to be quite a range of torus opening angles and the small error presented here on the mean angle should not be interpreted as the dispersion in opening angles present in the population. We have repeated the analysis of this section excluding CSO sources (projected linear sizes $`30`$ kpc). We find that this reduces the quasar fraction by $`0.04`$ in all the bins. The quasar fraction of all luminous CSO sources is $`0.56\pm 0.08`$, different at the 2$`\sigma `$ level from that of non-CSO sources. We defer discussion of possible reasons for a higher quasar fraction for CSOs in low-frequency selected samples to a future paper. ## 4 Possible causes of the change in quasar fraction with luminosity ### 4.1 Selection effects caused by lower intrinsic broad line luminosity We first investigate whether the decrease in quasar fraction with decreasing narrow-line and radio luminosity is a simple selection effect. If the low-luminosity objects have lower luminosity narrow emission lines it is natural to expect that their broad line and continuum luminosities should be lower too (e.g. Miller et al. 1992). It follows that their spectra may on average have lower signal-to-noise and weak broad lines may be missed. Given the strong correlation between narrow-line and radio luminosities (e.g. Willott et al. 1999) this could explain the deficit of broad line objects at radio luminosities $`\mathrm{log}_{10}(L_{151}`$ / WHz<sup>-1</sup>sr$`{}_{}{}^{1})<26.5`$, and also the apparent constancy of the quasar fraction for objects with luminous narrow lines discussed in Section 3. To test whether selection effects such as these could cause the apparent lack of low-luminosity broad line objects, we first calculate the expected broad line fluxes of objects in the 3CRR and 7C samples as a function of luminosity. To do this we make use of the fact that the narrow emission line luminosities of radio galaxies and quasars are positively correlated with low-frequency radio luminosity as shown in Willott et al. (1999). This correlation has a slope of $`0.79\pm 0.04`$ in the sense that $`L_{[\mathrm{OII}]}L_{151}^{0.79}`$ and is most likely due to a correlation between jet power and photoionizing continuum luminosity (Rawlings & Saunders 1991). Use of this relation enables one to estimate the narrow line \[OII\] luminosity for values of radio luminosity. A similar correlation between the broad line luminosity and radio luminosity is expected to hold. There are several lines of evidence supporting this. First, radio luminosity is correlated with the optical continuum luminosity for steep-spectrum quasars (Serjeant et al. 1998) and quasars show little variation in broad line equivalent widths over a wide range of luminosity (e.g. Osmer & Shields 1999). Second, Celotti et al. (1997) have shown that broad line luminosities are correlated with the power in pc-scale jets in radio-loud quasars. To determine the relationship between the strengths of narrow and broad lines in quasar/BLRG spectra, we adopt the line ratios from the composite quasar spectrum of Francis et al. (1991). Although this is from an optically-selected sample, we do not expect significant differences between the line ratios of radio-loud and radio-quiet quasars. This is borne out by a comparison with the Molonglo Quasar Sample of Baker & Hunstead (1995): repeating the analysis using their line ratios gives a virtually identical result. We consider a range of radio luminosities and attach to each one typical redshifts at which sources of these luminosities are observed in both the 3CRR and 7C samples. The correlation between the \[OII\] line and radio luminosity is then used to determine typical \[OII\] luminosities at each value of radio luminosity. From the typical wavelength range covered by optical spectra (4000Å– 8500Å), we find the strongest broad line observable at each redshift. The typical luminosities, redshifts and strongest broad lines are shown in Table 1. Finally, we use the ratio of broad to \[OII\] lines in Francis et al. (1991) to calculate the luminosity and hence flux of the strongest broad line expected at each luminosity-redshift pair. There is considerable scatter in the relationship between narrow emission line luminosity and radio luminosity ($`\sigma =0.5`$ dex). Extra scatter comes in due to the conversion from narrow to broad line luminosities. Therefore a total scatter of $`1\sigma =0.6`$ dex was assumed to determine the fraction of objects with broad line fluxes significantly below the characteristic values. Figure 2 shows the expected flux of the strongest broad line observable at each redshift as listed in Table 1 for both 3CRR and 7C. Also plotted are the actual data for the 7C-I and 7C-II quasars and BLRGs (Willott et al. 1998a) and all 3CRR quasars with $`z<0.86`$ in the RA range 18–12 hr (all from Jackson & Browne 1991 except 3C 109 from Goodrich & Cohen 1992). In general points corresponding to 3CRR BLRGs are not plotted because only a few of their broad line fluxes are available in the literature, although a few points from Hill et al. (1996) are plotted to illustrate the discussion of Section 4.2. The quasars which are red(dened) \[$`\alpha _{\mathrm{opt}}>1`$, where $`f_\nu \nu ^{\alpha _{\mathrm{o}pt}}`$\] are shown as open symbols on Fig. 2; again the 3CRR spectrophotometric dataset is too inhomogeneous for this to be done reliably for 3CRR sources. Note that very few quasars fall below the limits of the model error bars. Virtually all of those which do are reddened (the exception is 5C6.282 at $`z2`$, which has an unusual spectrum with very narrow Ly$`\alpha `$; see Willott et al. 1998a). The similarity of the predicted fluxes with those observed justifies the use of the line ratios and method described here. The key point to note from Fig. 2 is that as one considers lower radio luminosities in each sample (i.e. lower redshifts), the expected broad line fluxes increase. Therefore, so long as the spectra of lower redshift sources have similar exposure times to those at higher redshift, the broad lines should be visible if they are there and unreddened. For the 3CRR sample the spectra are rather inhomogeneous and it is difficult to prove that this is the case (but note that the deficit of 3CRR sources is most marked at $`z<0.3`$, where the H$`\alpha `$ line is covered by optical spectra and the expected broad H$`\alpha `$ fluxes are high). However, in the 7C sample, similar exposure times have been attained for the spectra at low redshift to those at high redshift. Therefore we are confident that in the absence of significant reddening of the BLR, an insignificant number of sources would be mis-classified. A further problem is whether a weak quasar spectrum can be discerned against a stronger host galaxy spectrum. To determine if this could cause broad line selection problems, synthetic spectra of quasars and galaxies were created and combined. The model used for the galaxy spectra was a 1 Gyr old stellar population synthesis model of Bruzual & Charlot (1993). For the quasar spectra, the LBQS composite of Francis et al. (1991) was used. Poisson noise was added and it was found that broad H$`\alpha `$ or MgII should still be clearly visible for a quasar $`2`$ magnitudes (in $`B`$-band) fainter than the galaxy (Fig. 3). Broad H$`\beta `$, however, is more difficult to detect. Spectra with poor blue wavelength coverage (e.g. no data below $`5000`$ Å) of sources at $`0.3<z<0.8`$ would not include H$`\alpha `$ or MgII, in which case H$`\beta `$ is the brightest line. Almost all the 7C and 6C spectra cover a large enough wavelength range to avoid this problem. However many spectra of 3CRR sources in the literature have a smaller wavelength range. But the radio–optical correlation (and Table 1) show that quasars at these redshifts in 3CRR typically have high luminosities and should not be fainter (at $`B`$-band) than their host galaxies. Indeed, the deficit of broad line objects in 3CRR only really begins at $`z\stackrel{<}{}0.3`$. Hence quasar broad lines should be clearly visible against typical background galaxy spectra down to $`\mathrm{log}_{10}(L_{151}`$ / WHz<sup>-1</sup>sr$`{}_{}{}^{1})25.5`$, with possible problems only at fainter radio luminosities. Another possibility to consider is that the narrow lines may be much stronger than the broad lines, so only weak wings in the line profiles would be seen. However, this picture seems unlikely for many sources in the absence of broad line reddening, since if both the narrow and broad lines are photoionizied by the same nuclear source it would require a conspiracy between broad and narrow line covering factors. We conclude that the low quasar fraction at low luminosity is not due to selection effects caused by the difficulty of detecting intrinsically less luminous broad lines. ### 4.2 Increased reddening at low luminosities In this section we consider the possibility that the drop in quasar fraction at low luminosities is due, at least in part, to an increase in the fraction of lightly-reddened ($`0\stackrel{<}{}A_V\stackrel{<}{}5`$) lines-of-sight as quasar luminosity decreases. It can be seen from Fig. 2 that the majority of the 7C quasars are lightly reddened at the highest redshifts. We exclude from this paper any discussion of any systematic changes of quasar reddening with cosmic epoch. The question of the range of reddening towards the nuclei of powerful (3CRR) $`z1`$ radio sources has recently been addressed by the $`3.5\mu `$m imaging programmes of Simpson, Rawlings & Lacy (1999) and Simpson & Rawlings (2000). They concluded that these high luminosity nuclei are mostly either naked ($`A_V0`$) or heavily obscured ($`A_V\stackrel{>}{}15`$) with only a small fraction ($`15`$ per cent) of intermediate lightly-reddened cases. If these results could be extrapolated to the lower narrow-line and radio luminosity regimes then light reddening of quasar nuclei would not be a serious issue for our study of the quasar fraction. However, two studies suggest that the fraction of lightly-reddened lines-of-sight might increase significantly at lower narrow-line and radio luminosities. First, Hill, Goodrich & DePoy (1996) studied the Pa$`\alpha `$ and $`H\alpha `$ lines in a complete sample of low redshift ($`0.1<z<0.2`$) 3CR sources, looking for evidence of reddened broad lines in radio galaxies. These objects lie near the top of lowest-$`L_{151}`$ (and $`z`$) bin in Fig. 1. They found broad lines in six objects out of a complete sample of thirteen (including two FRI sources), of which three had intermediate values of $`A_V`$. This study, although subject to a major problem with small number statistics together with the difficulties of detecting lightly reddened broad lines over certain ranges of redshift (see Fig. 2), hints that an increase in the fraction of the lightly-reddened population at low luminosities might provide part of the explanation for the dropping quasar fraction at low luminosities. Second, Baker (1997) showed that a typical low-luminosity lobe-dominated quasar, i.e. an object akin to those near the top of lowest-$`L_{151}`$ (and $`z`$) bin in Fig. 1, has a reddening $`A_V3`$ and is thus lightly reddened. We conclude that some of the drop in the quasar fraction at low luminosities might be explained by an increased chance of light reddening towards low luminosity quasar nuclei. In other words, we cannot allay the suspicion that some lightly reddened quasars have been misclassified as galaxies on the basis of existing, chiefly optical, spectroscopy. In the lowest-$`L_{151}`$ and $`z`$ bin of Fig. 1 this is a problem for (very low-$`z`$) 3CRR sources because of the inhomogenous optical dataset; it is a problem for the 6C and 7C sources because their redshifts are sufficient to remove the strong H$`\alpha `$ line from the optical window (see Fig. 2). To pursue this idea further we next consider specific physical models which might explain this behaviour. ### 4.3 The receding torus model Lawrence (1991) proposed that the apparent increase in quasar fraction and hence $`\theta _{\mathrm{trans}}`$ with luminosity in the 3CRR sample could be explained naturally by a ‘receding torus’ model. Dust sublimates at $`T1500`$ K and therefore the inner radius of a dusty torus/disc may be governed by the radius at which this temperature is achieved via radiation from the AGN. The more luminous the AGN the larger this radius and if the scale height of the torus $`h`$ is independent of luminosity, then more luminous sources will have a larger $`\theta _{\mathrm{trans}}`$; the inner radius of the torus $`r`$ scales as $`L^{0.5}`$ and $`\theta _{\mathrm{trans}}=\mathrm{tan}^1(r/h)`$. This model can also explain more lightly-reddened quasars at low luminosity since a larger fraction of the lines-of-sight pass through intermediate columns of obscuring material when the inner wall of the torus lies closer to the nucleus (see e.g. Fig. 8 of Hill et al. 1996). Simpson (1998) has shown that if the height of the torus is independent of the ionizing luminosity $`L`$ then the receding torus model predicts a theoretical quasar fraction $`f_\mathrm{q}`$ given by $$f_\mathrm{q}=1\left(1+\frac{L}{L_0}\mathrm{tan}^2\theta _0\right)^{0.5},$$ (1) where $`L_0`$ and $`\theta _0`$ are the normalisation luminosity and the torus half-opening angle at that luminosity, both of which are fixed by the measurement of the quasar fraction at some normalisation luminosity. We further assume that the \[OII\] luminosity scales with the ionizing continuum luminosity, and that it is not significantly affected by the changing torus opening angle, so that we can replace ionising luminosities in Eqn 1 by values of $`L_{[\mathrm{OII}]}`$. <sup>3</sup><sup>3</sup>3 As emphasised by Simpson, \[OII\] is probably a far from ideal choice of narrow emission line for this procedure because for an ionisation-bounded emission line cloud, the \[OII\] emission is from a region whose size depends much less strongly on the incident photoionising flux than the sizes of the regions responsible for higher-ionization lines like \[OIII\]. This leads to a much weaker dependence of \[OII\] luminosity on incident photoionising luminosity $`L`$ than the near proportionality between, say, \[OIII\] luminosity and $`L`$. Unfortunately, the choice of \[OII\] was forced on us by the available spectrophotometric data (see Willott et al. 1999). Binning all FRII sources in terms of $`L_{[\mathrm{OII}]}`$ we find that for $`35<\mathrm{log}_{10}(L_{[\mathrm{OII}]}/\mathrm{W})<36`$, the quasar fraction is $`0.32\pm 0.04`$ giving $`\theta _0=46^{}`$. This is used to normalise Equation 1 which is plotted in Fig. 4 along with the quasar fraction measured in four bins of $`\mathrm{log}_{10}(L_{[\mathrm{OII}]})`$. This simple model appears to fit the data fairly well although it does predict a marginally sharper transition between low and high quasar fractions than the data (at about the $`1\sigma `$ level). Note that the above model depends crucially upon the height of the torus being independent of luminosity. If the height of the torus increases with luminosity, albeit more slowly than the inner radius, then this would flatten the model distribution to something more closely resembling the data. Taken together with the results of Hill et al. (1996), we conclude that the receding torus model provides a plausible if non-unique explanation for the declining quasar fraction below $`\mathrm{log}_{10}(L_{[OII]}/\mathrm{W})=35.1`$. ### 4.4 Dust–ejection by jets or outflow Tadhunter et al. (1999) have recently obtained high-resolution near-infrared imaging of the powerful radio galaxy Cygnus A which shows an edge-brightened biconical structure roughly aligned with the jets. They suggest that at least some of the very high reddening towards the nucleus of Cygnus A ($`A_V70`$ mag) occurs in a kiloparsec scale dust-lane. For radio-loud sources, the jets may clear a path through this dust, so that only when viewed within a critical angle of the jet-axis is there a chance of a line-of-sight having little reddening. The jet power could control the dust-clearing efficiency in such a way as to give an anti-correlation between reddening and radio luminosity. The conceptual difference between this model and the receding torus model is that it is now a decrease in jet power (and hence the radio luminosity) rather than a decrease in optical luminosity that controls the closing down of the effective opening angle of the radiation cone at low luminosities. Tadhunter et al. speculate that lightly reddened quasars are objects viewed within the nuclear torus opening angle, but not within the region cleared of dust by the jets. Another possibility (e.g. Dopita et al. 1998) is that dust is cleared by a radiation-driven outflow in which the opening angle and strength of the outflow could be related to the luminosity of the nucleus and hence give rise to the observed dependence of quasar fraction on luminosity. ### 4.5 Dual-population models The evidence of increased reddening at low luminosities (Section 4.2) and the quantitative success of the receding torus model (Section 4.3) hint that we have a reasonable explanation for at least part of the drop in quasar fraction at low luminosities. In this section we discuss separate evidence for a second conceptually different mechanism. A corollary of Fig. 4 is that at low luminosities the low fraction of quasars results from a small value for the typical opening angle of the torus $`\theta _{\mathrm{trans}}`$. Falcke et al. (1995) have cited this as a strength of the receding torus model since if the opening angle approaches the beaming angle of the jet then there will be an unobscured view of the nucleus of low power jets only when their emission is also strongly Doppler-boosted. Such objects would be classified as BL Lac objects rather than quasars, providing a neat explanation for the well-known lack of quasars with FRI radio structures. We are unaware of any systematic measurements of broad line fluxes in BL Lac type objects which would provide a quantitative test of this hypothesis. However, recent observations of BL Lacertae itself (Corbett et al. 2000), an object with an extended radio flux characteristic of a low-luminosity radio galaxy (Antonucci & Ulvestad 1985), suggest that in at least this object there are broad emission lines at about the predicted level, and that these lines are excited by a hot accretion disc. However, other recent results suggest that the small opening angles required at low luminosities from Fig. 4 are more likely to argue against the generality of the receding torus model. If it were true that low luminosity radio galaxies possessed very thick obscuring tori then it should prove very difficult to get a clear view of their central regions at optical wavelengths. Chiaberge, Capetti & Celotti (1999) have studied the nuclear emission from FRI galaxies using the HST. In most cases they find probable optical synchrotron from the nuclear regions, which they interpret as implying a clear view of the inner jet and hence a lack of obscuring material on scales greater than $`100`$ pc. This straightforwardly rules out a kpc-scale obscuration model for FRIs along the lines of the model proposed by Tadhunter et al. (1999) for Cyg A. Still stronger constraints come from the nearby FRI radio galaxy M87 since its optical synchrotron core is smaller than $`5`$ pc and varies on the timescales of months (Tsvetanov, Kriss & Ford 1997), implying that the emission region is a pc or smaller in size. Since the inner regions of the optical synchrotron-emitting jet are in clear view, and these size scales are comparable in size to the expected size of the obscuring torus in low-luminosity quasars (e.g. Netzer & Laor 1993), these results begin to question the existence of any thick torus, at least in the case of M87. Chiaberge et al. maker stronger statements than this by making the additional assumption that this optical emission is emitted from within the region probed by radio VLBI which is argued to be within $`0.1`$ pc of the central black hole (Junor, Biretta & Livio 1999). We remain to be convinced that these observations rule out the existence of a pc-scaled torus in M87. There is a plethora of evidence from studies of M87 using optical continuum (Chiaberge et al. 1999), emission-line (Ford et al. 1994) and hard X-ray (Guainazzi & Molendi 1999) observations that it contains a supermassive black hole accreting at a tiny ($`<10^4`$) fraction of its Eddington-limited rate (see Willott et al. 1999 for a discussion of this object in the context of the radio-optical correlation). As shown in Fig. 1, excluding objects with weak narrow emission lines gives a quasar fraction which is virtually independent of redshift and/or radio luminosity. This observation can be naturally explained if the weak-line objects belong to a different population from the strong emission line objects. Perhaps, as is seemingly the case for M87, a large fraction of these objects are fundamentally different in that they lack a well-fed quasar nucleus and also a thick obscuring torus (Chiaberge et al. 1999), and perhaps their jets are powered by a fundamentally different mechanism (e.g. Blandford & Begelman 1999). The continuity of the radio-optical correlation for radio sources provides a weak argument against a fundamental change in the accretion process (Rawlings & Saunders 1991), but evidence for systematic changes in the slope of this relation at low luminosities (Zirbel & Baum 1995) may be evidence that there are at least some differences at low radio luminosities. Arguing along similar lines, Laing et al. (1994) proposed that the low-excitation radio galaxies (LEGs) in their 3CRR sample would lack broad lines if viewed from any orientation. Recall that we specifically excluded FRI radio sources from our analysis of the quasar fraction in Section 3. The only difference that including the FRIs would have made, would be to decrease the quasar fraction in the lowest luminosity bin from $`0.14\pm 0.04`$ to $`0.10\pm 0.03`$. Such a ‘two population’ model would remove another troubling problem for any extrapolation of the receding torus model to the lowest quasar luminosities. Seyfert galaxies – radio-quiet AGN with similar optical narrow-line luminosities to radio galaxies in the lowest-$`L_{151}`$ bin of Fig. 4 – do not have the narrow opening angles predicted by the receding torus model. Direct observations of the ionization cones in nearby Seyfert galaxies typically show cones with half-opening angles in the range $`2050^{}`$ (Wilson & Tsvetanov 1994), considerably greater than those predicted. Moreover, the vast majority of Seyfert galaxies do not seem to be obscured, i.e. of type 2, as would be required by narrow opening angles. Hence, as noted by Falcke et al (1995) it would appear that the receding torus model cannot apply similarly to radio-quiet and radio-loud AGN requiring some ad-hoc explanation, for example a different geometry for the obscuring material (i.e. a smaller scale height), due to a difference in black hole mass, environment and/or angular momentum. We can make a crude association of the putative second population of radio sources with all sources with narrow emission line luminosities $`\mathrm{log}_{10}(L_{[\mathrm{OII}]}/\mathrm{W})<35.1`$. Due to the scatter in the radio–optical correlation, it is not expected that this would lead to a clean division between the two populations at a particular \[OII\] luminosity, and if the data were available, a classification based on line ratios, and hence excitation, might well prove cleaner. However, subject to this limitation, the change in quasar fraction with redshift of a combination of the two populations (seen on the left plot of Fig. 1) is then naturally explained by the less rapid cosmic evolution of the low-luminosity population (e.g. Urry & Padovani 1995) than the high-luminosity population (shown on the right plot of Fig. 1). Hence at high-redshift, the contribution of the low-luminosity population is negligible and the quasar fraction of 0.4 is just that of the high-luminosity population. Dual population modelling of the low-frequency radio luminosity function is consistent with just such a scheme (Jackson & Wall 1999; Willott et al. in prep.). Also, Laing et al. (1994) and Hardcastle et al. (1998) find that low-excitation radio galaxies have linear size and core prominence distributions consistent with an isotropic population. There are two residual concerns with the two population model. First, many of the low-luminosity objects have emission lines, so their excitation and the correlation between luminosity and ionization parameter (Saunders et al. 1989; Tadhunter et al. 1998) need some explanation. Second, it must explain why radio galaxies and quasars have different narrow line luminosity distributions at intermediate luminosities (Jackson & Browne 1990) but similar distributions at high luminosities (Jackson & Rawlings 1997). The first concern can be addressed by simple analogy with M87: the absence of a broad-line quasar nucleus does not mean an absence of a photoionising source; for example Dopita et al. (1997) favour radiative shocks as a source of the excitation for the H$`\alpha `$ lines in the nuclear disc of M87, and shock models in which ionization parameter correlates with line luminosity are easily envisaged (e.g. Dopita & Sutherland 1995). The second concern is probably also easily dealt with. Considering first the 3CRR sample, the dual-population model is consistent with the drop in quasar fraction at low luminosities because a low luminosity 3CRR source is necessarily at low redshift where there is clearly a mixture of both populations; at high redshifts only the high-luminosity population is observed. Thus any comparative narrow emission line study of quasars and radio galaxies which is based on bright radio samples (Jackson & Browne 1990; Jackson & Rawlings 1997) should yield different line luminosity distributions at low redshift, and similar distributions at high redshift, which is just as observed. However, a small difference in the distributions of the \[OII\] line luminosities of intermediate luminosity quasars and radio galaxies is also observed in the 7C sample (Willott et al. 1999). The lower radio flux limit of this sample means that these intermediate luminosity sources are at high redshift, where there are few low emission line luminosity objects, so the difference is not due to the mixing of the two populations. This is most likely the result of small but inevitable biases pointed out by Rawlings & Saunders (1991), and quantified in the context of the receding torus model by Simpson (1998): a positive correlation between quasar luminosity and opening angle, coupled with inevitable scatter in quasar luminosity at a fixed radio luminosity, means that objects viewed within the opening angle are biased towards the more luminous objects within the scatter. ### 4.6 Time variability There is yet one more possible cause of the drop in quasar fraction at low luminosities: systematic differences in the time variability of objects with luminosity. In Willott et al. (1999) we used the narrow emission line–radio correlation to suggest that the most luminous objects are probably accreting at rates close to the Eddington limit, but lower luminosity sources are sub-Eddington accreters. It therefore seems plausible that the lower luminosity sources have more scope for variability, since an object accreting at the Eddington rate should have a ready fuel supply which is accreted at a fairly steady rate limited by radiation pressure. In contrast, sub-Eddington accretion suggests there is not a ready supply of fuel available and hence fluctuations in accretion rate may be more likely. Indeed observations of radio-quiet quasars show that the lower luminosity quasars are more highly optically-variable than higher luminosity quasars over timescales of a few years (e.g. Véron & Hawkins 1995; Cristiani et al. 1996; Paltani & Courvoisier 1997), although it should be noted that some authors attribute at least some of this variability to gravitational microlensing (e.g. Hawkins & Taylor 1997). The small quasar fraction at low luminosities could be explained by these objects spending a significant fraction of their active lifetimes in a ‘quiet’ state whereas the more luminous objects are continuously active over their entire lifetime. Due to light travel time effects, different emission regions of quasars have different variability timescales: BLR $``$ months; NLR $`10^4`$ yr; radio lobes $`10^6`$ yr. Reverberation mapping of quasars has shown that the broad line fluxes follow the nuclear continuum flux with just such a time lag (see Peterson 1993 for a review). If a quasar undergoes high-amplitude variability over a timescale $`100`$ yr, then only the continuum and BLR fluxes would be observed to undergo this strong variability, and the NLR and extended radio emission would simply reflect the time-averaged output of the central engine. Note that a quasar undergoing a decrease in luminosity of this sort of timescale may then appear as a low-excitation radio galaxy. If this behaviour occurs only in a significant fraction of the intermediate and low luminosity objects, then it could reproduce the observed change in quasar fraction. Essentially this model is a slightly more complicated version of the two-population model considered in Section 4.5, with the two populations being unified along a time rather than a jet orientation axis. It is important to stress that there is as yet no firm evidence for the type of time variability required. However, the few long-term time variability studies hint that there may be some interesting effects: the BLRG 3C 390.3 was found to undergo a sustained decrease of 1.5 mags in the optical over $`80`$ years (Cannon, Penston & Penston 1968); however, Angione (1973) used archival plates to study the variability of 23 quasars over $`60`$ years finding no evidence for long-term sustained increases or decreases. A separate issue related to time variability is whether the probability of individual radio sources being observed to be quasars varies systematically throughout their lifetimes. Because flux-limited samples introduce inevitable biases into the age distributions of the sources they contain (e.g. Blundell, Rawlings & Willott 1999), this can lead to subtle effects. ## 5 Conclusions Using complete, low-frequency selected samples of radio-loud AGN we have investigated the fraction of observed broad line objects – the quasar fraction – as a function of redshift, and radio and emission line luminosity. Our findings are interpreted in terms of orientation-based unified schemes. We find that * Considering only those sources with strong narrow emission lines \[$`\mathrm{log}_{10}(L_{[\mathrm{OII}]}/\mathrm{W})35.1`$, corresponding roughly to $`\mathrm{log}_{10}(L_{151}/\mathrm{W})26.5`$\], the quasar fraction is virtually independent of radio luminosity and redshift, and is consistent with a simple unified scheme involving an obscuring torus with a half-opening angle $`\theta _{\mathrm{trans}}`$ of $`53^{}`$. * For less luminous emission-line sources, the quasar fraction is much lower, a finding which is not simply the result of selection effects induced by the lower intrinsic broad line luminosities of these sources. We have found evidence which supports at least two probable physical causes for the drop in quasar fraction at low luminosity: (i) a gradual decrease in $`\theta _{\mathrm{trans}}`$ and/or a gradual increase in the fraction of lightly-reddened ($`0\stackrel{<}{}A_V\stackrel{<}{}5`$) lines-of-sight with decreasing quasar luminosity; and (ii) the emergence of a second population of low luminosity radio sources which, like M87, lack a well-fed quasar nucleus, and may well lack a thick obscuring torus. ## Acknowledgements We would like to thank Steve Eales, Gary Hill, Julia Riley and David Rossitter for important contributions to the 7C Redshift Survey. Thanks also to Robert Laing and Chris Simpson for some very useful discussions, and to the referee Ian Browne for useful suggestions. This research has made use of the NASA/IPAC Extra-galactic Database, which is operated by the Jet Propulsion Laboratory, Caltech, under contract with the National Aeronautics and Space Administration. CJW thanks PPARC for support.
warning/0003/cond-mat0003504.html
ar5iv
text
# Possible scenario of the spatially separated Fermi-Bose mixture in the superconductive bismuthates Ba1-xKxBiO3 ## I Introduction The concept of the extremely strong-coupling superconductivity with the preexisted local pairs was firstly introduced by Shafroth in the middle of fifties. His statement was that in the extremely type-II superconductors, where the parameter $`\xi _0k_F1`$, the nature of the superconductive transition corresponds to the local pair formation (pairing in the real, and not in the momentum space) at some relatively high temperature $`T^{}`$, and their Bose-Einstein condensation (BEC) at a lower critical temperature $`T_c<T^{}`$. Later on Alexandrov and Ranninger developed this concept in connection with the narrow-band materials with an extremely strong electron-phonon coupling constant ($`\lambda 1`$), where a standard Eliashberg theory starts to fail. The key issue of their approach was a statement that in narrow bands, where polaron formation is important, it is possible, in principle, to create the conditions where the two polarons could effectively occupy the same potential well, prepared in a self-consistent fashion. Approximately in the same time Leggett and Nozieres developed a general theory which yields a smooth interpolation between a BCS-type of pairing in the momentum space for a small electron-electron attraction and the pairing in the real space for a large electron-electron attraction. There were two crucial points in the papers . (i) Their results are valid independently of the precise nature of the short-range effective attraction between electrons; (ii) they investigated self-consistently a standard equation for the superconductive gap and an equation for the number of particles conservation. The most important result of Nozieres and Leggett is that for $`T^{}>T_c`$ (or, in other words, for a binding energy of a local pair $`|E_b|>\epsilon _F`$) the chemical potential $`\mu `$ is always large and negative $`\mu =|E_b|/2+\epsilon _F<0`$. Hence in a strong-coupling limit a superconductive pairing takes place not on the Fermi-surface, as in the BCS-theory, but below the bottom of a conductive band. This is a crucial drawback of all local-pairs theories. We cannot match two basic facts: the existence of the Fermi-surface and the presence of preformed pairs above $`T_c`$. First one who emphasized this contradiction was J. Ranninger , who introduced the concept of the two-band Fermi-Bose mixture. In this scenario the presence of a degenerate fermionic band guarantees the existence of a Fermi-surface, while the Bose-Einstein condensation is responsible for the superconductivity in a bosonic band. Soon after the discovery of high-$`T_c`$ superconductors P. W. Anderson reintroduced the concept of the local pairs. He also introduced two bands of the fermionic and bosonic quasiparticles. In his approach, a superconductive transition was connected with the BEC in the bosonic band of the charge excitations — holons, while the presence of a large Fermi-surface was guaranteed by the fermionic band of the spin excitations — spinons. Unfortunately, even this beautiful approach was not totally successful because at least in a one layer the BEC of holons yields a charge of a superconductive pair equal to $`e`$ instead of $`2e`$ measured experimentally. Later on Geshkenbein, Ioffe, and Larkin phenomenologically reintroduced a concept of the Fermi-Bose mixture on the level of the coefficients in the Ginzburg-Landau expansion and showed that several important experiments in the underdoped high-$`T_c`$ materials can be explained naturally within this form of the Ginzburg-Landau functional. So, up to now a question whether a two-band Fermi-Bose mixture scenario is applicable to the high-temperature superconductive (HTSC) materials is still open. Probably, the best chances to be described by this scenario has a bismuth family of high-$`T_c`$ superconductors Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub>, where the parameter $`k_F\xi _02`$, and the tunneling experiments of Fischer et al. signal the formation of a pretty large and a stable pseudogap at temperatures well above $`T_c`$. In our paper we would like to discuss a possibility of a two-band Fermi-Bose mixture scenario in a quite different class of superconductors with a relatively high $`T_c30`$ K, namely in the superconductive bismuthates Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub> (BKBO). The key issue of our paper is a possibility of the existence of the two spatially separated bands of fermionic and bosonic quasiparticles in these materials. Our paper is organized as follows. In the first part we present the basic experimental facts concerning the local electronic and crystal structure peculiarities, and their connection with the superconductive and the normal transport properties of BKBO. In the second part we try to show how these basic facts could be naturally explained within a scenario of the two spatially separated bands of the fermionic and bosonic quasiparticles. We conclude the paper by a summary of our model and a discussion of the several additional experiments, which would help us to give a definite answer whether our proposal is the only possibility for a superconductive pairing in the bismuthates. ## II Spatially separated Fermi-Bose mixture The BaBiO<sub>3</sub>, which is a parent compound for the bismuthates Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub> and BaPb<sub>1-x</sub>Bi<sub>x</sub>O<sub>3</sub> (BPBO), represents a charge-density-wave (CDW) insulator having the two gaps: an optical gap $`E_g=1.9`$ eV and an activation (transport) gap $`E_a=0.24`$ eV . A partial replacing of Ba by K in BKBO causes the decrease of the gaps, and as a result the insulator-metal transition as well as the superconductivity are observed at the doping levels $`x0.37`$. The superconductivity remains up to the doping level $`x=0.5`$ corresponding to the solubility limit of K in BKBO but a maximal critical temperature $`T_c30`$ K is achieved for $`x0.4`$ . ### A Local crystal structure peculiarities A three-dimensional character of a cubic perovskite-like structure of the bismuthates differs from a two-dimensional one in the HTSC cuprates. The building block in the bismuthates is a BiO<sub>6</sub> octahedral complex (analogue of CuO<sub>n</sub> ($`n=4,5,6`$) in the HTSC-materials). The octahedral complexes are the most tightly bound items of the structure because of a strong covalence of the Bi$`6s`$-O$`2p_\sigma `$ bonds. According to the crystallographic data , the crystal structure of a parent BaBiO<sub>3</sub> represents the alternating arrangement of the expanded and contracted BiO<sub>6</sub> octahedra in a barium lattice, spoken as a “breathing” distortion. Such an alternation together with a static rotation of the octahedra around axis produce the monoclinic distortion of a cubic lattice. As shown in Refs. , to the larger soft octahedron corresponds the complex BiO<sub>6</sub> with the completely filled Bi$`6s`$-O$`2p`$ orbitals and to the smaller rigid octahedron corresponds a complex BiL<sup>2</sup>O<sub>6</sub>. Here L<sup>2</sup> denotes the free level in the antibonding Bi$`6s`$-O$`2p_\sigma ^{}`$ orbital of the smaller octahedral complex. The K doping of BaBiO<sub>3</sub> is equivalent to a hole doping and leads to a partial replacement of the larger soft octahedra BiO<sub>6</sub> by the smaller rigid octahedra BiL<sup>2</sup>O<sub>6</sub> . This causes a decrease and finally a disappearance of the static breathing and tilting distortions, and the lattice should contract despite a practically equal ionic radii of K<sup>+</sup> and Ba<sup>2+</sup>. As a result, the average structure according to the neutron diffraction data at the doping level $`x=0.37`$ becomes a simple cubic. However, the local EXAFS probes showed the essential difference of the local crystal structure from the average one. We found out that the oxygen ions belonging to the different BiO<sub>6</sub> and BiL<sup>2</sup>O<sub>6</sub> octahedra vibrate in a double-well potential, while those having equivalent environment of the two equal BiL<sup>2</sup>O<sub>6</sub> octahedra oscillate in a simple harmonic potential . This very unusual behavior is in a close connection with a local electronic structure of BKBO. ### B Local electronic structure The coexistence in BaBiO<sub>3</sub> of the different types of the octahedra with the two Bi-O bond lengths and strengths reflects the different electronic structures of BiO<sub>6</sub> complexes. The valence band of BaBiO<sub>3</sub> is determined by the overlap of Bi$`6s`$ and O$`2p`$ orbitals , and, owing to a strong Bi$`6s`$-O$`2p_\sigma `$ hybridization, the octahedra can be considered as the quasi-molecular complexes . In each complex there are ten electron levels consisting of the four bonding-antibonding Bi$`6s`$-O$`2p_\sigma `$ orbitals and the six nonbonding O$`2p_\pi `$ orbitals. A monoclinic unit cell includes the two octahedra and contains 38 valence electrons (10 from the two bismuth ions, 4 from the two barium ions, and 24 from the six oxygen ions). All the Bi-O bond lengths should be equal and the local magnetic moments should be present in the case of equal electron filling for nearest octahedra (BiL<sup>1</sup>O$`{}_{6}{}^{}+`$BiL<sup>1</sup>O<sub>6</sub>). In contrast both the presence of the two types of octahedral complexes and an absence of any local magnetic moment are experimentally observed , so the mentioned above scheme of the valence disproportionation 2BiL<sup>1</sup>O$`{}_{6}{}^{}`$ BiL<sup>2</sup>O<sub>6</sub>+BiO<sub>6</sub> was proposed . In this scheme the numbers of occupied states in the neighbouring octahedral complexes are different: the octahedron BiL<sup>2</sup>O<sub>6</sub> contains 18 electrons and has one free level or a hole pair L<sup>2</sup> in the upper antibonding Bi$`6s`$O$`2p_\sigma ^{}`$ orbital, while in the octahedron BiO<sub>6</sub> with 20 electrons this antibonding orbital is filled as shown in Fig. 1. It is quite natural that the BiL<sup>2</sup>O<sub>6</sub> octahedra have the stiff (quasi-molecular) Bi-O bonds and a smaller radius, while the BiO<sub>6</sub> octahedra represent the non-stable molecules with a filled antibonding orbital and a larger radius. Because the sum of the two nearest octahedra radii overcomes the lattice parameter, the octahedral system must tilt around aixis, producing together with a breathing a monoclinic distortion in BaBiO<sub>3</sub>. Thus, in BaBiO<sub>3</sub> one has an alternating system of the two types of the octahedral complexes filled with the local pairs: the hole pairs in BiL<sup>2</sup>O<sub>6</sub> complexes and the electron pairs in BiO<sub>6</sub> complexes. The local pair formation in BaBiO<sub>3</sub> was advocated previously, see for example . The binding mechanism for the pairs is probably of an electronic nature in accordance with Varma’s picture of the pairing due to the skipping of the valence “4+” by the Bi ion . However one cannot fully exclude the lattice mediated pairing in accordance with de Jongh’s statement that the preference to retain the closed-shell structures overcomes the Coulomb repulsion involved with an intrasite bipolaron formation. The local electronic structure of BaBiO<sub>3</sub> combined with the real-space local crystal structure is presented in Fig. 2(a). In such a system there are no free Fermi carriers, and the conductivity occurs only due to the transfer of the carrier pairs . The dissociation of the pairs and the hopping of a single electron from one octahedron to another, in similarity with Varma’s suggestion, cost an energy: $$E_b=2E(\mathrm{Bi}\underset{¯}{\mathrm{L}}^1\mathrm{O}_6)[E(\mathrm{BiO}_6)+E(\mathrm{Bi}\underset{¯}{\mathrm{L}}^2\mathrm{O}_6)],$$ and is observed experimentally as an optical conductivity peak at the photon energy $`h\nu =1.9`$ eV . Thus we have an example of the normal bosonic liquid of the pairs bound with an energy $`E_b`$ (as in Ref. ). Experimentally, BaBiO<sub>3</sub> shows a semiconductor-like behavior with an energy gap $`E_a=0.24`$ eV, which can be explained only as a two-particle transport with the activation energy $`2E_a`$ due to the delocalization of the pairs. From our point of view, the transport gap is defined by the combined effect of the Coulomb interaction and the deformation potential between the neighboring octahedral complexes. In principle, the electronic structure of BaBiO<sub>3</sub> presented in Fig. 3(a) is similar to that proposed by Namatame et al. . They supposed that the thermally excited charge carriers in BaBiO<sub>3</sub> are the polarons with a substantial band narrowing due to a strong electron-phonon coupling. In our case the carriers are the local pairs (the bipolarons), the empty and the occupied bosonic bands are very narrow, and that is why in Fig. 2(a) they are shown as energy levels. So, the transport gap $`2E_a`$ is the gap between the empty and the occupied bosonic states, and the conductivity in BaBiO<sub>3</sub> arises from the thermo-activated bosonic quasiparticles. This fact is in agreement with the suggestion of Taraphder et al. about a bosonic character of the ground state in BaBiO<sub>3</sub>. Another key question in the bismuthates is a nature of the optical gap $`E_g=1.9`$ eV. In accordance with the band structure calculations , this gap arises from the peak-to-peak splitting of the Bi$`6s`$ band due to CDW formation. Namatame et al. attribute this gap to the Frank-Condon transition under which the lattice should be frozen. That contradicts to the observation in the Raman spectra of the abnormally large amplitude of the breathing-type vibrations of the oxygen octahedra. The required mode arises only at a resonant excitation by the laser radiation with an energy $`h\nu =E_g`$. The resonance was destroyed and the abrupt decrease of the mode amplitude was observed when the lasers with an other frequency were used . Obviously, the photons cannot excite a carrier pair as a whole and should destroy it. Such a behavior directly proves that the lattice is involved in this optical transition. In our scheme an excitation over the optical gap corresponds to the pair destruction BiL<sup>2</sup>O<sub>6</sub>+BiO$`{}_{6}{}^{}\stackrel{h\nu }{}`$ 2BiL<sup>1</sup>O<sub>6</sub>, which produces a local deformation of the lattice due to the changing of the two different octahedra on equivalent ones. This dynamic deformation is manifested in the Raman spectra as an abnormally high amplitude of the breathing mode. Thus the nature of the optical gap is just the pair binding energy $`E_g=E_b`$. It is important to emphasize that there are no free fermions in the system. Only the excited fermions can be produced by the unpairing, and they do not give any input into the charge transport because of a high value of $`E_b=1.9`$ eV. Thus the bosonic and the fermionic subsystems are separated both spatially and energetically, and therefore the Fermi-Bose mixture is absent in the parent compound. ### C Formation of the Fermi-Bose mixture The substitution of the each two K<sup>+</sup> for the two Ba<sup>2+</sup> modifies the BiO<sub>6</sub> complex to the BiL<sup>2</sup>O<sub>6</sub> one. As a result, the number of the small stiff BiL<sup>2</sup>O<sub>6</sub> octahedra increases as $`n_0(1+x)/2`$, while the number of large soft BiO<sub>6</sub> octahedra decreases as $`n_0(1x)/2`$, where $`n_0=1/a^3`$ is the number of the unit cells and $`a`$ is a lattice parameter. The spatial overlap of the BiL<sup>2</sup>O<sub>6</sub> complexes appears at the finite doping levels, which, taking into account their small radii and the rigid bonds, contracts the lattice. The structural changes are accompanied by the essential changes in the local electronic structure and in the physical properties of BKBO. A spatial overlap of the L<sup>2</sup> levels leads to their splitting into an empty fermionic-like band $`F`$ inside the BiL<sup>2</sup>O$`{}_{6}{}^{}\mathrm{}`$BiL<sup>2</sup>O<sub>6</sub> Fermi-cluster \[see Fig. 2(b)\]. In the doping range $`x<0.37`$ the band is narrow enough due to a polaronic effect and is still separated from an occupied Bi$`6s`$O$`2p`$ subband. The number of the empty electronic states in the $`F`$ band increases with $`x`$ as $`\widehat{n}_F=n_0(1+x)`$, while the number of the local electron pairs decreases as $`n_B=n_0(1x)/2`$. A free motion of the pairs is still prevented by the intersite Coulomb repulsion , which becomes strongly screened inside the clusters. When the Fermi-clusters are formed, the conductivity occurs due to the motion of the pairs through the clusters of the different lengths. The BKBO compounds demonstrate a semiconducting-like conductivity changing from a simple activation type to the variable-range-hopping Mott’s law . Moreover the activation energy lowers with the doping down to $`E_a0`$ at $`x0.37`$. One can understand the decrease of the activation energy as a suppression of the Coulomb blockade due to the formation of the Fermi-clusters and due to the decrease of an energy shift between the empty and the occupied bosonic levels as the lattice distortion is diminished. At the doping level $`x0.37`$ (see Fig.2(c) and Fig.3(c)) the following cardinal changes take place: (i) Both the breathing and the rotational static lattice distortions transform to the dynamic ones. At the cluster borders, where all oxygen ions belongs to BiO<sub>6</sub> and BiL<sup>2</sup>O<sub>6</sub> octahedra, the local breathing dynamic distortion is observed as a vibration in a double-well potential of $`(1x)`$ part of the oxygen ions but cannot be detected by the integral methods of the structural analysis such as an X-ray and a neutron diffraction. (ii) The infinite percolating Fermi cluster (formed from the spatially overlapped BiL<sup>2</sup>O<sub>6</sub> octahedra) appears, which leads to the overlap of an empty fermionic band with a filled one, and as a result $`F`$ becomes a conduction band. Overcoming of the percolation threshold provides the insulator-metal phase transition and the formation of the Fermi-liquid state for $`x>0.37`$. The valence electrons of the BiL<sup>2</sup>O<sub>6</sub> complexes previously localized become itinerant which is in agreement with the experiments . (iii) The pair localization energy disappears $`E_a0`$ so the local electron pairs (from the BiO<sub>6</sub> complexes) can freely move in the real space providing a bosonic contribution into the conductivity. Thus, in the metallic phase the two types of carriers are present: the itinerant electrons from the BiL<sup>2</sup>O<sub>6</sub> complexes (fermions) and the delocalized electron pairs from the BiO<sub>6</sub> complexes (bosons). Despite the normal state conductivity is mainly due to a fermionic subsystem, the contribution from a bosonic subsystem was also observed by Hellman and Hartford as the two-particle normal state tunneling. As a result at doping levels $`x>0.37`$ we have a new type of a spatially separated mixture of the bosonic $`B`$ and the fermionic $`F`$ subsystems, which describes metallic properties of BKBO. It should be stressed that the fermions and the bosons belong to the complexes with the different electronic structure, therefore the Fermi and the Bose subsystems are spatially separated at any doping level. These subsystems are connected by the relations $`2n_B+\widehat{n}_F=2n_0`$ and $`2n_B/\widehat{n}_F=(1x)/(1+x)`$. The high enough value of the binding energy, which in the superconductive compositions becomes apparent as a pseudo-gap $`E_b=E_g0.5`$ eV , is the guarantee against the pair destruction. The unpairing is possible only under the optical excitation to the band $`F^{}`$ (see Fig. 3), which does not play any role in the charge transport. At $`x=1`$ all the BiO<sub>6</sub> octahedra are transformed to the BiL<sup>2</sup>O<sub>6</sub> ones. The Bose system disappears ($`n_B=0`$) together with an excited fermionic band $`F^{}`$. As a result, KBiO<sub>3</sub> should behave as a simple Fermi-liquid metal without superconductive properties (see Fig.2(d) and Fig.3(d)). It is worth to notice that a metallic KBiO<sub>3</sub> compound exists only hypothetically because of the exceeding of the potassium solubility limit in BKBO, which in this case equals to $`x0.5`$. However BaPbO<sub>3</sub> as an electronic analogue of KBiO<sub>3</sub> demonstrates the metallic but not the superconductive properties. Recent attempts to synthesize KBiO<sub>3</sub> at a high pressure found out that only K<sub>1-y</sub>Bi<sub>y</sub>BiO<sub>3</sub> with a partial replacement of K<sup>+</sup> ions by Bi<sup>3+</sup> ones is formed . Such a replacement should lead to the appearance of the BiO<sub>6</sub> octahedra with the local electron pairs, and hence the compound K<sub>1-y</sub>Bi<sub>y</sub>BiO<sub>3</sub> should be superconductive in accordance with the discussion above. Indeed, a superconductivity with $`T_c=10.2`$ K was experimentally observed in this compound . From this point of view, it follows that BaPbO<sub>3</sub> should be superconductive at a partial substitution of the Ba<sup>2+</sup> ions for the La<sup>3+</sup> ones (or for the other trivalent ions) because such a substitution produces the local electronic pairs as in the case of K<sub>1-y</sub>Bi<sub>y</sub>BiO<sub>3</sub>. ## III Discussion The most important point which differs our model from the ones considered by Rice and Sneddon , Varma , and De Jongh is the claim that the local hole pair belongs to the whole BiL<sup>2</sup>O<sub>6</sub> complex and not only to the Bi<sup>5+</sup> site. Analogously the local electron pair belongs to the whole BiO<sub>6</sub> complex and not only to the Bi<sup>3+</sup> site. From this point of view we do not believe in the scheme 2Bi$`{}_{}{}^{4+}`$ Bi<sup>3+</sup>+Bi<sup>5+</sup> of a charge disproportionation, but propose a scheme 2BiL<sup>1</sup>O$`{}_{6}{}^{}`$ BiL<sup>2</sup>O<sub>6</sub>+BiO<sub>6</sub>, where L denotes a local hole in the upper antibonding Bi$`6s`$O$`2p_\sigma ^{}`$ orbital of the whole octahedral complex. To some extent our idea has a certain similarity with Zhang-Rice construction for HTSC-materials . The essential difference is that Zhang-Rice singlet is a boson (a holon in Anderson terminology) with zero spin and charge $`e`$ (so to create a Cooper pair one needs two singlets). In the bismuthates from the beginning we have a rather tightly bound boson (biholon) with a proper charge $`2e`$. Of course, a total spin of the biholon is again zero, so their Bose-Einstein condensation corresponds to a standard $`s`$-wave superconductivity. The existence of the biholons before was only proved rigorously in the quasi one-dimensional ladder materials at a strong coupling along the rungs (Dagotto and Rice ). Our scheme immediately gives us an understanding of a spatial separation of the fermionic and the bosonic bands in the bismuthates. Namely, our model identifies an optical gap with the binding energy of a preformed pair $`E_b=E_g`$. In the same time a transport gap $`2E_a`$ corresponds to the pair localization energy. As a result we have the picture for the one-particle density of states presented in Fig. 3. In this picture for $`x=0`$ there is a filled bosonic band separated by the large gap $`E_g=E_b`$ from an empty fermionic band (an excited band $`F^{}`$) and by a smaller transport gap $`2E_a`$ from an empty bosonic band $`B`$, which plays the role of the conduction band for the bosonic quasiparticles involved in the activation transport. In accordance with Ref. in the representation of the one-particle density of states the filled bosonic band has the hole-like dispersion while the empty bosonic band has the electron-like dispersion. The above bands are on top of the completely filled wide ($``$16 eV) Bi$`6s`$-O$`2p`$ hybridized band, which includes 18 valence electrons per unit cell. Since the occupied and the empty bands belong to the different octahedra, they are always spatially separated. Formally, the transport gap is similar to the impurity band in doped semiconductors. However, the combined transport and the Hall-effect measurements showed that the number of carriers participated in an activation conductivity is extraordinary large and is estimated as a number of unit cells . This rejects any impurity mechanisms and suggests that the conduction process in the doping range $`0<x<0.37`$ can not be understood by the conventional transport mechanisms of the ordinary semiconductors. The attempts have been made to identify the excitation with the energy gap $`E_a`$ using some optical methods: the reflectivity, the photoconductivity, and the photoacoustic spectroscopy measurements. However, no photoresponse could be detected by either method in the energy region near $`E_a`$ . So the meaning of this gap is just a localisation energy of the local pairs. Hence for $`x=0`$ we have the normal bosonic semiconductor with an activation character of the conductivity $`\sigma (T)=\sigma _0\mathrm{exp}(2E_a/kT)`$. As we pointed above, the localization energy is influenced by at least two effects: the Coulomb intersite repulsion due to CDW state and the energy shift between the empty and the occupied bosonic bands due to a static lattice distortion. With the doping both effects diminish, which leads to the decrease of the pair localization energy. The reason is that the appearing Fermi clusters screen the Coulomb intersite repulsion and the decrease of the static lattice distortion lowers the energy shift between the bosonic bands. At $`0<x<0.37`$ the lengths of the Fermi-cluster islands along directions are variable due to the random distribution of the dopant atoms, so the temperature dependence of the conductivity becomes more complicated, but remains a semiconducting-like. It corresponds to the experimentally observed variable-range-hopping conductivity according to the Mott’s law $`\sigma =\sigma _0\mathrm{exp}(T_0/T)^{1/4}`$ . The value of $`T_0(36)\times 10^8`$ K obtained by Hellman et al. implies a strong carrier localization as in our model. On the language of the fermionic and the bosonic bands, at $`0<x<0.37`$ the Fermi and the Bose subsystems are both energetically and spatially separated. Near the insulator-metal transition the characteristic temperature $`T_0`$ depends on the stoichiometry, and an agreement with the Mott’s law breaks down . A pair localization energy approaches zero because the movement of BiO<sub>6</sub> octahedron, surrounded by the equivalent BiL<sup>2</sup>O<sub>6</sub> octahedra in undistorted lattice, does not change the total energy. In the doping range $`x0.37`$ we have a two-band Fermi-Bose mixture similar to that proposed earlier by Ranninger et al. . In the same time this mixture of a bosonic metal state with a Fermi-liquid state is rather unusual. The normal conductivity in this mixture is mainly due to the fermionic subsystem which overcomes at $`x0.37`$ the percolation threshold. Nevertheless, the bosonic two-particle contribution into the normal state conductivity is also present. At the doping level $`x0.37`$ the occupied and the empty pair levels are in the resonance, therefore the energy separation disappears and the Bose and the Fermi subsystems remain separated only spatially. Note that there is an interplay between the Bose and the Fermi subsystems since the motion of the carrier pairs leads to the transformation of the Bose octahedral clusters to the Fermi ones and vice versa in the process of the dynamic exchange BiO$`{}_{6}{}^{}\mathrm{Bi}`$L<sup>2</sup>O<sub>6</sub>. Because this process is closely related with the superconductivity, we analyze it below in more details. ### Superconductivity in Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub> Taking into account an existence of the double-well potential in Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub> one can consider superconductivity in this compound in the framework of the anharmonic models for HTSC . As it was shown in these models, if the oxygen ions move in a double-well potential, an order-of-magnitude enhancement of a constant of an electron-lattice coupling follows automatically from a consistent treatment of this motion. The pairing mechanism is connected in these models with the enhancement of the coupling constant due to the oxygen ion vibrations in a double-well potential. However, in agreement with Refs. , we believe that in the bismuthates the pairing mechanism is more probably of the electronic than of the phonon-mediated origin (see Sec. II b). In the same time in our system the local pairs tunnel between the neighboring octahedra due to the vibration of the oxygen ions in the double-well potential. Therefore we suppose that the lattice is involved in the superconductivity more probably by providing the motion of a local pair rather than via a pairing mechanism itself. The process of a dynamic exchange is illustrated in Fig. 4. An oxygen belonging to the two neighboring octahedra BiO<sub>6</sub> and BiL<sup>2</sup>O<sub>6</sub> vibrates in a double-well potential, and hence the tunneling of the electron pair between the neighboring octahedra occurs when the ion tunnels through the potential barrier between the wells. Because of this interconnection between the processes of the pair and the oxygen tunneling, one can estimate the matrix element of the pair tunneling as $`t_B\omega _0e^D`$ where $`\omega _0`$ is the tunneling frequency, $`D=(1/\mathrm{})_{x_0}^{x_1}|p|𝑑x(d/\mathrm{})\sqrt{2MU}`$ is a quasiclassical transparency of the barrier in the double-well potential, $`U`$ and $`d`$ are the barrier height and width, and $`M`$ is the oxygen ion mass. Note that rather small tunneling frequency $`\omega _0=200`$ K (see Fig.4) already incorporates the effects connected with a polaronic narrowing of the one-particle bands $`t_pt_0exp(g^2)`$ and the hopping of a bosonic pair via virtual dissociation processes. The last ones are described by a second order perturbation theory (see Ref. ) and yield an estimate $`t_Bt_p^2/|E_b|`$ for a width of a bosonic band in a spatially homogeneous Bose-system. The tunneling of the pairs helps to establish a macroscopic long range order (a phase coherence) in the bosonic system. On the language of the spatially separated Fermi-Bose mixture, a local pair is transferred from one Bose cluster to a nearest one over the Fermi-cluster, which, depending on the doping level, consists of several octahedra. The pairs overcome the Fermi-cluster step by step. A single step, corresponding to the pair transfer into a neighboring octahedron, is described by the pair tunneling in the double-well potential. Thus the tunneling frequency $`\omega _0`$ is the same for each step. If one assumes that the steps are independent events, a probability of the overcoming of the Fermi-cluster can be obtained as a product of the probabilities of an each step. In this case the matrix element of the pair tunneling through the Fermi-cluster can be estimated as $`\widehat{t}_B\omega _0e^{ND}`$ where an evarage number of steps (which is proportional to a Fermi-cluster linear size) can be obtained from the ratio of the concentrations of BiL<sup>2</sup>O<sub>6</sub> and BiO<sub>6</sub> octahedra. Hence a number of steps can be estimated as $`N\left(\frac{1+x}{1x}\right)^{1/3}`$. Of course, it is naturally to assume that the critical temperature of superconductivity is of the order of the temperature of the Bose–Einstein condensation $`T_c\widehat{t}_Ba^2n_B^{2/3}`$ in the bosonic system with a large effective mass $`m_B1/\widehat{t}_Ba^2`$. We remind that $`a^3n_B=(1x)/2`$ in our case. For $`x=0.4`$ and the parameters of the double-well potential obtained in the Ref. (see also Fig. 4) we estimated $`T_c50`$ K. This value is larger than the measured $`T_c30`$ in the bismuthates. Note that the estimate above is quite rough. An accurate analysis of the superconductivity in the bismuthates should also take into account a significant boson-phonon interaction, arising due to the interconnection between the vibrations of the oxygen ions and the transfer of the pairs. When the pair is transferred from one octahedron to another, the lattice has a sufficient time to relax, forming each time a new configuration before the next tunneling event occurs. As a result the pair’s “deformed” environment (the BiO<sub>6</sub> octahedra) may follow the tunneling processes without the retardation. If the local pair motion is slow compared to the frequencies of the optical phonons associated with the deformations of the octahedra, the so-called anti-adiabatic limit is fulfilled in our system . Note that in our case the stretching phonons are associated with the tunneling of the pairs. The following conclusions can be made from the analysis of the phonon modes studied in Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub> by an inelastic neutron scattering . (i) The frequencies of the optical modes of the Bi-O vibrations $`\omega _{ph}>630`$ K are high enough to provide the anti-adiabatic limit, since the tunneling frequencies are lower, $`\omega _0300`$ K for $`x0.5`$ . Note that from a theoretical point of view a polaronic narrowing of the one- particle bandwidths $`t_p<<t_0`$, and hence a small value of $`\omega _0`$ itself, is just a direct consequence of the non-adiabaticity. (ii) The energy of the longitudinal stretching mode with wave-vector direction is lower than the energy of modes with other wave-vector directions or of transversal modes. (iii) The breathing-type vibrations with the wave vector $`𝐪_b`$=$`(\pi /2a,0,0)`$ are energetically favorable since an energy of the longitudinal stretching phonons is the lowest at the Brillouin band edge. That is why a “breathing” of each octahedron should be coordinated with its neighbors to guarantee a resonant tunneling in the system. Hence a long-range correlation of the vibrations should occur at low temperatures when only the low-energy states are occupied. (iv) The bandwidth of the longitudinal stretching mode is of the order of 100 K, and thus a temperature $`TT_c`$ is high enough to excite the non-breathing-type longitudinal stretching phonons with the wave vectors shorter than $`𝐪_b`$. The thermal excitation of the phonons with such short wave-vectors leads to the destruction of the long-range correlation between the breathing-type vibrations, and hence play a destructive role for the pair tunneling. It should be stressed in addition, that the oscillations of the oxygen ion in the double-well potential corresponds to the breathing-type longitudinal vibrations with directions (see Fig. 4a). Thus the pair motion is more correctly described as follows. When the local pair is transferred to a neighboring octahedron (for example, from the left to the right in Fig. 4c) due to the transition of the oxygen ion from one well to another one in the double-well potential, the charge density corresponding to a pair is quickly redistributed inside the octahedron. As a result a double-well potential, designed for the vibration of the other oxygen ions belonging to the same octahedron, is formed. The longitudinal stretching phonons with $`\omega _{ph}>\omega _0`$ are involved in this fast process. After the relaxation of the deformed surrounding, the pair becomes prepared for the next hopping, and the described process repeats again and again, providing a resonant tunneling of the pairs on the large distances along directions. Note that since the BiO<sub>6</sub> and BiL<sup>2</sup>O<sub>6</sub> complexes have different strengths of the Bi-O bonds, the pair transfer in turn is able to change both the phase and the wave vectors of the stretching phonons. As the vibration in the double-well potential is of a breathing-type, the wave vectors remain unchanged when the octahedra ( which the pair is passing by) vibrates in the breathing mode. However a value of $`𝐪`$ can be changed in the case of the vibrations with the short wave-vector. Due to the dispersion of the longitudinal stretching mode such a change can cost some energy. Thus to provide the pair motion in this case, an energy should be transmitted to the phonons and the dissipation of the pair kinetic energy should occur. Of course, this leads to the decrease of the BEC critical temperature for the pair. Hence the more exact estimate for $`T_c`$ requires the solution of a problem of a self-consistent preparation of the barrier due to the interaction of a pair with a phonon subsystem in a process of underbarrier tunneling. This brings us all the nice physics of the tunneling with the dissipation . Remind, that in the problem of the tunneling with the dissipation the shape and the height of a potential barrier in a two-level system are determined self-consistently taking into account an interaction of a tunneling particle with a thermal reservoir. Note that the more exact evaluation of $`T_c`$ will give us an experimentally observed decrease of the critical temperature as a function of the concentration in the metallic region $`x>0.4`$. This decrease should take place mainly due to the following facts: (i) a decrease of a bosonic density $`n_B`$, (ii) an increase of the width of the barrier $`N`$, and maybe the most important, (iii) the decrease of lattice softening which leads to the increase of the dissipation of the pair kinetic energy due to the stretching phonons with $`𝐪𝐪_b`$. Note also, that at temperatures $`T>T_c`$ a bosonic subsystem behaves for the concentrations $`x>0,37`$ as a normal bosonic metal with a heavy mass $`m_B1/\widehat{t}_Ba^2`$. Summarizing the discussion above we point out that the two processes are important for the superconductivity in Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub>. The vibrations of the oxygen ions in the double-well potential provides the mechanism for a transfer of the local pairs from the one Bose-cluster to the other. At the same time the pair motion is strongly affected by the stretching longitudinal vibrations of the oxygen ions in the octahedra which the pair is passing by. We suppose that the last process should be taken into account to estimate correctly the critical temperature. It is worth to notice, that a similar dispersion of the longitudinal stretching phonons, which leads to the dominant role of the breathing-type phonons at low temperatures, has been observed also for the high $`T_c`$ cuprates La<sub>1.85</sub>Sr<sub>0.15</sub>CuO<sub>4</sub> and YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> . Taking into account the recent experimental evidence by Müller et al. of the coexistence of the small bosonic and fermionic charge carriers in La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub>, we suppose also to apply our Fermi-Bose mixture scenario to HTSC cuprates. Our paper on this problem is in preparation . Note again that the nature of the pairing itself in the cuprates is definitely of the electronic origin. However, due to the fact that underdoped HTSC-materials are close to the phase-separation on AFM and PM-clusters , the lattice here can play an assistant role again providing a pair tunneling between the superconductive PM metallic clusters via an insulating AFM-barrier. It can also serve as a limitation on the estimate of the effective critical temperature for the Bose-condensation of the pairs in our system. ## IV Conclusion In conclusion we briefly summarize the main results. 1. The parent compound BaBiO<sub>3</sub> represents a system with the initially preformed local electron and hole pairs. Every pair is spatially and energetically localized inside an octahedron volume. The localization energy of a pair determines the transport activation gap $`E_a`$. The binding energy of a pair becomes apparent as the optical gap $`E_g=E_b`$. 2. The new type of the Fermi-Bose mixture is possibly realized in the superconductive compositions of Ba<sub>1-x</sub>K<sub>x</sub>BiO<sub>3</sub> for $`x>0.37`$. The bosonic bands are responsible for the two-particle normal state conductivity. The overlap of the empty fermionic band $`F`$ with an occupied valence band Bi6$`s`$O2$`p`$ provides the insulator-metal phase transition and produces the Fermi-liquid state. This state shunts to a great extent the normal state conductivity arising from the two-particle Bose transport. 3. The fermionic band $`F^{}`$ connected with the pair destruction does not play any role in the transport. The excitation energy is high enough to guarantee against the destruction of bosons (a pair binding energy $`E_b0.5`$ eV for the superconductive compositions). 4. Pair localization energy is absent for $`x>0.37`$ ($`E_a=0`$), so the bosonic and the fermionic subsystems are separated only spatially. The interplay between them is due to the dynamic exchange BiL<sup>2</sup>O$`{}_{6}{}^{}`$BiO<sub>6</sub>, which causes a free motion of the local pairs in the real space. 5. The pairing mechanism in the bismuthates is more probably of the electronic than of the phonon-mediated origin. The existence of the local pairs and their tunneling between the neighboring octahedra are the reasons for the appearance of the double-well potential, which describes the vibration of the oxygen ions. The lattice is involved in the superconductivity more probably by providing the motion of the local pairs. Finally, we would like to emphasize that the scenario of the Fermi-Bose mixture allows us to describe qualitatively an insulator-metal phase transition and a superconductive state in BKBO in the framework of the one common approach. To some extent this scenario explains the contradictions observed experimentally by the UPS and XPS , the EXAFS and XANES , and the Raman spectroscopies, as well as by an inelastic neutron scattering , the transport and the optical measurements . Nevertheless, the additional experiments are required to make a definite conclusion about the nature of the superconductivity in these systems. First of all we propose two direct experiments to test our model. (i) To synthesize a new compound Ba<sub>1-x</sub>La<sub>x</sub>PbO<sub>3</sub>, which should be superconductive in accordance with our point of view. (ii) To provide the Raman scattering experiment of the superconducting Ba<sub>0.6</sub>K<sub>0.4</sub>BiO<sub>3</sub> compound using a resonance optical excitation in the range of the optical pseudogap $`E_g0.5`$ eV. In this case the appearance of the additional Raman modes due to local dynamic distortions should be observed at the pair destruction in accordance with our model. Besides, it is important to carry out the more precise measurements of the specific heat in the bismuthates for $`TT_c`$. We know that the specific heat behaves as $`C_B(T/T_c)^{3/2}`$ for the temperatures $`T<T_c`$, and $`C_B=const`$ for $`TT_c`$ in a three-dimensional Bose-gas. As a result, there is a $`\lambda `$-point behavior of the specific heat for $`TT_c`$. However, in the Fermi-Bose mixture there is an additional contribution from a degenerate Fermi-gas $`C_F\gamma T`$. This contribution could in principle destroy a $`\lambda `$-point behavior of the specific heat in the Fermi-Bose mixture. Note that the currently available experimental results in the bismuthates signal a smooth behavior of the specific heat near $`T_c`$ , because in all the experiments the contributions from the degenerate Fermi and Bose-gases are masked by a larger lattice contribution. Another important measurement, which can be proposed to elucidate the nature of the superconductivity, is a measurement of the thermopower in the normal state of the bismuthates. According to the ideas of Larkin et al. , the Seebeck coefficient $`S_0=n_B(T)/n_0\mathrm{ln}(n_B(T)/n_0)+\alpha T/\epsilon _F`$ in the Fermi-Bose mixture is much larger by absolute value than the usual value $`S_0=\alpha T/\epsilon _F`$ in the normal state of an ordinary fermionic metal. Moreover, the sign of the thermopower can become negative, which is also rather unusual. Note that up to now there is a lot of controversy in the measurements of Seebeck coefficient in the bismuthates (see Ref. and references therein). ###### Acknowledgements. We acknowledge fruitful discussion with N. M. Plakida, Yu. Kagan, P. Fulde, P. Woelfe, and A. N. Mitin. This work was supported by Russian Foundation for Basic Research (Grants 99-02-17343 and 98-02-17077), INTAS grant 97-0963, and Program “Superconductivity” (Grant 99010). M. Yu. K is grateful for the grant of the President of Russia 96-15-96942 and acknowledges a hospitality of Max-Planck-Institut in Dresden.
warning/0003/astro-ph0003376.html
ar5iv
text
# CROSS-CORRELATING CMB POLARIZATION WITH LOCAL LARGE SCALE STRUCTURES ## 1 Introduction Secondary CMB anisotropies offer new windows to constrain cosmological models. Lens effects$`^\mathrm{?}`$ are particularly attractive since they are expected to be one of the dominant effect. Methods to detect the lens effects on CMB$`^\mathrm{?}`$ have been proposed recently. Unfortunately all of them suffer from a high sensitivity to cosmic variance. However, this problem can be solved if one considers CMB polarization instead of temperature anisotropies. Standard cosmological models predict that at small scale, the so called $`B`$ component of the polarization can be significant only if CMB-lens couplings are present$`^{\mathrm{?},\mathrm{?}}`$. This feature of CMB polarization will allow us to present tools that enhance the detection of lens effect in CMB by mixing them with galaxy survey $`^\mathrm{?}`$. ## 2 Lens effects on CMB polarization Photons emerging from the last scattering surface are deflected by the large scale structures of the Universe that are present on the line-of-sights. Therefore photons observed from apparent direction $`\stackrel{}{\alpha }`$ must have left the last scattering surface from a slightly different direction, $`\stackrel{}{\alpha }+\stackrel{}{\xi }(\stackrel{}{\alpha })`$. Thus, the lensing effect does not produce any polarization nor rotate the polarization vector, it just moves the apparent direction of the line of sight$`^\mathrm{?}`$; if $`\stackrel{}{P}=(Q,U)`$ the Stoke variables, $$\stackrel{}{\widehat{P}}(\stackrel{}{\alpha })=\stackrel{}{P}(\stackrel{}{\alpha }+\stackrel{}{\xi }).$$ (1) This mechanism alters the geometrical properties of the polarization field, that is to say, changes the *electric* ($`E`$) and *magnetic* ($`B`$) components of the polarization that reflects its non-local geometrical properties. Indeed, inflationary models, predict that the small scale $`B`$ polarization is dominated by lens effect $`^{\mathrm{?},\mathrm{?}}`$. We explicit this point in the weak lensing regime where distortions are small. At leading order (one can refer to $`^\mathrm{?}`$ for description of this calculation) one obtains: $$\mathrm{\Delta }\widehat{B}=2ϵ_{ij}\left(\gamma ^i\mathrm{\Delta }\widehat{P}^j+\gamma _{,k}^i\widehat{P}^{j,k}\right)$$ (2) where we described the lens effect by its convergence field $`\kappa =1/2\xi _{,i}^i`$ and its shear field $`(\gamma _1,\gamma _2)=1/2(\xi _{,x}^x\xi _{,y}^y,2\xi _{,x}^y)`$. The $`ϵ_{ij}`$ (the totally antisymmetric tensor) reflects the geometrical properties of the $`B`$ field. It comes in front of two shear-polarization mixing terms. One which we will call the $`\mathrm{\Delta }`$-term couples the shear with second derivative of the polarization field. The other one, hereafter the $``$-term, mixes gradient of the shear and polarization. As a consequence, the $`B`$ field directly reflects the properties of the shear maps. Fig. 1 shows a comparison of relation (2) with the exact lensing effect. The agreement is excellent. ## 3 Cross-correlating CMB maps and weak lensing surveys With the help of eq.(2), one can try to recover lensing information out of $`B`$ polarization. Unfortunately, a direct inversion is not possible since it leads to a huge degeneracy in the resulting shear maps $`^\mathrm{?}`$. Another way of deciphering the encoded lensing data will be to cross-correlate CMB polarization with other lensing information, namely, weak lensing galaxy surveys$`^\mathrm{?}`$. There are strong theoretical motivations to perform this kind of cross-correlations$`^{\mathrm{?},\mathrm{?}}`$. The cross-correlation coefficient between line-of-sight mapped by a photon emerging from last scattering, $$r=\frac{\kappa _{\mathrm{cmb}}\kappa _{\mathrm{gal}}}{\sqrt{\kappa _{\mathrm{cmb}}^2\kappa _{\mathrm{gal}}^2}},$$ (3) is around 40% in standard models and accordingly, the correlation between $`B`$ polarization and lensing survey will be significant. Since only the lens effect generates $`B`$ field at this scale, we can expect to have a low cosmic variance on the cross-correlation. Moreover, one can assume that systematics and foreground noises that will hamper each detections will be poorly correlated, so that mixing the two data sets can be an effective way of enhancing the accuracy of the signal. We present here two objects that mix CMB polarization data with reconstructed shear fields. Looking at eq. (2), the most simple idea is to try to construct *guess* $`B`$-fields, $`b_\mathrm{\Delta }`$ and $`b_{}`$, with local shear instead of the CMB one and to try to correlate them with our polarization data. $`b_\mathrm{\Delta }ϵ_{ij}\gamma _{\mathrm{gal}}^i\mathrm{\Delta }\widehat{P}^j,b_{}ϵ_{ij}_k\gamma _{\mathrm{gal}}^i_k\widehat{P}^j.`$ (4) Then, the amplitude of the cross-correlation between $`\mathrm{\Delta }B`$ and $`b_\mathrm{\Delta }`$ can easily be estimated. At leading order, we have $$\mathrm{\Delta }\widehat{B}b_\mathrm{\Delta }(\stackrel{}{\alpha })=\mathrm{\Delta }E^2\kappa \kappa _{\mathrm{gal}},\mathrm{\Delta }\widehat{B}b_{}(\stackrel{}{\alpha })=\frac{1}{2}(\stackrel{}{}E)^2\stackrel{}{}\kappa \stackrel{}{}\kappa _{\mathrm{gal}}.$$ (5) These results remain valid even when filtering effects are included (see $`^\mathrm{?}`$ for complete calculation). Fig. 1 shows a numerical illustration of this cross-correlation. The similarities between the left and the right maps are not striking. Yet, under close examination one can recognize individual patterns shared between the maps. Moreover, computation of correlation coefficient gives significant overlapping up to 40%. Using this results, we define two quantities insensitive to the normalization of CMB and $`\sigma _8`$ and to filtering effects, which probe the cross-correlation between two lensing planes: $`𝒳_\mathrm{\Delta }{\displaystyle \frac{\mathrm{\Delta }\widehat{B}b_\mathrm{\Delta }(\stackrel{}{\alpha })}{\mathrm{\Delta }\widehat{E}^2\kappa _{\mathrm{gal}}^2}}={\displaystyle \frac{\kappa \kappa _{\mathrm{gal}}}{\kappa _{\mathrm{gal}}^2}},𝒳_{}{\displaystyle \frac{\mathrm{\Delta }\widehat{B}b_{}(\stackrel{}{\alpha })}{(\stackrel{}{}\widehat{E})^2(\stackrel{}{}\kappa _{\mathrm{gal}})^2}}={\displaystyle \frac{1}{2}}{\displaystyle \frac{\stackrel{}{}\kappa \stackrel{}{}\kappa _{\mathrm{gal}}}{\stackrel{}{}\kappa _{\mathrm{gal}}^2}}`$ (6) Previous methods to probe weak lensing in CMB anisotropies $`^{\mathrm{?},\mathrm{?}}`$ ran into high cosmic variance problems. This is not surprising since lens effect which is not dominant can be masked by statistical deviations of the primary CMB signal, thus reducing the accuracy of lens detection. Since $`B`$ polarization only emerges in presence of lensing, this last effect should be less important. Indeed, we showed in $`^\mathrm{?}`$ that $`𝒳_\mathrm{\Delta }`$ cosmic variance can be simply estimated in terms of the cosmic variances of the polarization field and of the shear field. $`\mathrm{C}osVar(𝒳_\mathrm{\Delta })=\mathrm{C}osVar\left(\mathrm{\Delta }E^2\right)+\left({\displaystyle \frac{1+r^2}{2r^2}}\right)\mathrm{C}osVar\left(\kappa ^2\right).`$ (7) The same kind of equation holds for $`𝒳_{}`$. This expression leads to values for the cosmic variance of $`𝒳_\mathrm{\Delta }`$ of less than 8% for realistic 100 square degrees surveys (see table 1). ## 4 Conclusion - Sensitivity to the cosmic parameters We showed that weak lensing effect on the CMB $`B`$ polarization can be embedded in a simple, real space, first order expression. This expression can be used to create mathematical objects that compare the lensing effects up to the last scattering surface to the one up to our galaxy surveys probes. We showed that this objects should produce a significant information, even in realistic (i.e. filtered) situations, with a low intrinsic statistical error. These objects are also expected to be good, unbiased, cosmic parameters tracers. Fig. 2 presents their behavior in the $`(\mathrm{\Omega }_0,\mathrm{\Lambda })`$ plane which exhibit a high sensitivity to the vacuum energy density. This is not surprising, since we are probing the length of the optical bench we are working in, which is rather sensitive to $`\mathrm{\Lambda }`$ $`^{\mathrm{?},\mathrm{?}}`$. ## Acknowledgments We thank B. Jain, U. Seljak and S. White for the use of their ray-tracing simulations. KB and FB thank CITA for hospitality and LvW is thankful to SPhT Saclay for hospitality. We are all grateful to the TERAPIX data center located at IAP for providing us computing facilities. ## References
warning/0003/hep-ph0003164.html
ar5iv
text
# VACUUM ENERGY: “IF NOT NOW, THEN WHEN?” ## 1 A Cosmological Constant or Quintessence? Absent a known symmetry principle protecting its value, no theoretical reason for making the cosmological constant zero or small has been found. Inflation makes the universe flat, so that, at present, the vacuum or smooth energy density $`\mathrm{\Omega }_{Q0}=1\mathrm{\Omega }_{m0}<1`$, is $`10^{120}`$ times smaller than would be expected on current particle theories. To explain this small but non-vanishing present value, a dynamic vacuum energy, quintessence, has been invoked, which obeys the equation of state $`w_QP/\rho <0`$. (The limiting case, $`w_Q=1`$, a static vacuum energy or Cosmological Constant, is homogeneous on all scales.) The evidence for a flat low-density universe come from : (1) The location of the first Doppler peak in the CBR anisotroy at $`l200`$: $`\mathrm{\Omega }_m+\mathrm{\Omega }_Q=1\pm 0.2`$; (2) The slow evolution of rich clusters, the mass power spectrum, the CBR anisotropy, the cosmic flow:$`\mathrm{\Omega }_{m0}=0.3\pm 0.05`$; (3) Curvature in the SNIa Hubble diagram, dynamic age, height of first Doppler peak, cluster evolution: $`\mathrm{\Omega }_{Q0}=1\mathrm{\Omega }_{m0}2/3`$. Of these, the SNIa evidence is most subject to systematic errors due to precursor intrinsic evolution and the possibilty of grey dust extinction. The combined data nevertheless implies a flat, low-density universe with $`\mathrm{\Omega }_{m0}1/3`$ and a smooth energy component with present energy density $`\mathrm{\Omega }_{Q0}2/3`$ and negative pressure $`1w_Q1/2`$. Accepting this small but non-vanishing value for static or dynamic vacuum energy, a flat Friedmann cosmology (CDM) is characterized by $`\mathrm{\Omega }_{m0},\mathrm{\Omega }_{Q0}=1\mathrm{\Omega }_{m0}`$ or the present ratio $$_0u_0^3\mathrm{\Omega }_{Q0}/\mathrm{\Omega }_{m0}=(1\mathrm{\Omega }_{m0})/\mathrm{\Omega }_{m0},$$ and by the equation of state for the smooth energy. The Cosmic Coincidence problem now becomes pressing: Why do we live when the clustered matter density $`\mathrm{\Omega }(a)`$, which is diluting as $`a^3`$ with cosmic scale $`a`$, is just now comparable to the static vacuum energy or present value of the smooth energy i.e. when the ratio $`_02`$ ? In this paper, we distinguish the two limiting cases allowed for the smooth energy component: LCDM: Cosmological constant: $`w_Q=1`$ and QCDM: Quintessence: $`w_Q=1/2`$ . In the next section, we compare the expansion of these two limiting low-density flat universes. In Section 3, we extend to QCDM the calculation of asymptotic mass fraction as function of a hypothetical continuous variable $`\mathrm{\Omega }_{m0}`$ presented by Martel et al for QCDM. Finally, we statistically infer that, absent any prior information about $`\mathrm{\Omega }_{m0}`$, the observed present ratio $`_0`$ is reasonable for a LCDM universe, and most likely for a QCDM universe: “If not now, then when?” ## 2 Expansion of a Low Density Flat Universe The Friedmann equation in a flat universe with clustered matter and smooth energy density is $$H^2(x)(\dot{a}/a)^2=(8\pi G/3)(\rho _m+\rho _Q),$$ or, in units of $`\rho _{cr}(x)=3H^2(x)/8\pi G`$, $`1=\mathrm{\Omega }_m(x)+\mathrm{\Omega }_Q(x),`$ where the reciprocal scale factor $`xa_0/a1+z\mathrm{}`$ in the far past, $`0`$ in the far future. With the effective equation of state $`wP/\rho =`$ constant, different kinds of energy density dilute at different rates $`\rho a^n,n3(1+w)`$, and contribute to the deceleration at different rates $`(1+3w)/2`$ shown in the table: The expansion rate in present Hubble units is $$H(x)/H_0=(\mathrm{\Omega }_{m0}x^3+(1\mathrm{\Omega }_{m0})x_Q^n)^{1/2}.$$ The Friedmann equation has an unstable fixed point in the far past and a stable attractor in the far future. (Note the tacit application of the anthropic principle: Why does our universe expand, rather than contract?) The second Friedmann equation is $`\ddot{a}a/\dot{a}^2=(1+3w_Q\mathrm{\Omega }_Q)/2`$. The ratio of smooth energy to matter energy, $`(a)=_0(a_0/a)^{3w_Q}`$, increases as the cosmic expansion dilutes the matter density. A flat universe, characterized by $`R_0,w_Q`$, evolves out of an SCDM universe in the remote past towards a flat de Sitter universe in the future. As shown by the inflection points (O) on the middle curves of Figure 1, for fixed $`_0`$, QCDM expands faster than LCDM, but begins accelerating only at the present epoch. The top and bottom curves refer respectively to a de Sitter universe ($`\mathrm{\Omega }_m=0`$), which is always accelerating, and an SCDM universe ($`\mathrm{\Omega }_m=1`$), which is always decelerating. The matter-smooth energy transition (“freeze-out”) $`\mathrm{\Omega }_Q/\mathrm{\Omega }_m=1`$ took place only recently at $`x^{w_Q}=_0^{1/3}u_0`$ or at $`x=1+z=u_0^2=1.59`$ for QCDM and, even later, at $`x=1+z=u_0=1.26`$ for LCDM. Because, for the same value of $`u_0`$, a matter-QCDM freeze-out would take place earlier and more slowly than a matter-LCDM freeze-out, it imposes a stronger constraint on structure evolution. As summarized in the table below, quintessence dominance begins 3.6 Gyr earlier and more gradually than cosmological constant dominance. (In this table, the deceleration $`q(x)\ddot{a}/aH_0^2`$ is measured in present Hubble units.) The recent lookback time $$H_0t_L(z)=z(1+q_0)z^2+\mathrm{},z<1,$$ where $`q_0=0`$ for QCDM and $`=1/2`$ for LCDM. ## 3 Evolution of Large Scale Structure In this section, we extend to QCDM earlier LCDM calculations of the asymptotic mass fraction $`f_{c,\mathrm{}}`$ that ultimately collapses into evolved galaxies. This is presumably a measure of the number density of galaxies like our own, that are potentially habitable by intelligent life. We then compare the QCDM and LCDM asymptotic mass fraction distribution functions, as function of an assumed $`\mathrm{\Omega }_{m0}`$. The background density for large-scale structure formation is overwhelmingly Cold Dark Matter (CDM), consisting of clustered matter $`\mathrm{\Omega }_m`$ and smooth energy or quintessence $`\mathrm{\Omega }_Q`$. Baryons, contributing only a fraction to $`\mathrm{\Omega }_m`$, collapse after the CDM and, particularly in small systems, produce the large overdensities that we see. Structure formation begins and ends with matter dominance, and is characterized by two scales: The horizon scale at the first cross-over, from radiation to matter dominance, determines the power spectrum $`P(k,a)`$, which is presently characterized by a shape factor $`\mathrm{\Gamma }_0=\mathrm{\Omega }_{m0}h=0.25\pm 0.05`$. The horizon scale at the second cross-over, from matter to smooth energy, determines a second scale factor, which for quintessence, is at $`130`$ Mpc, the scale of voids and superclusters. A cosmological constant is smooth at all scales. Quasars formed as far back as $`z5`$, galaxies at $`z6.7`$, ionizing sources at $`z=(1030)`$. The formation of any such structures, already sets an upper bound $`x<30`$ or $`(\mathrm{\Omega }_\mathrm{\Lambda }/\mathrm{\Omega }_{m0})<1000,\mathrm{\Omega }_{Q0}<30`$, for any structure to have formed. A much stronger upper bound, $`u_0<5`$, is set by when typical galaxies form i.e. by estimating the probability of our observing $`_0=2`$ at the present epoch. ### 3.1 Asymptotic Collapsed Mass Parameter $`\beta `$ Martel et al and Garriga et al have already calculated the asymptotic mass fraction from the Press-Schechter formalism $$f_{c,\mathrm{}}=\text{erfc}(\sqrt{\beta })=(2/\sqrt{\pi })_\sqrt{\beta }^{\mathrm{}}\mathrm{exp}(t^2)𝑑t,$$ depending only on $$\beta \delta _{i,c}^2/(2\sigma _i^2),$$ where $`\sigma _i^2`$ is the variance of the denssity field, smoothed on some scale $`R_G`$, and $`\delta _{i,c}`$ is the minimum density contrast at recombination which will ultimately make a bound structure. This minimum density contrast grows with scale factor $`a`$, and is, except for a numerical factor of order unity , $`\delta _{i,c}x/(1+z_i)`$. Both numerator and denominator in $`\beta `$ refer to the epoch of recombination, but this factor $`(1+z_i)`$ cancels out in the quotient. (MSW and MS have improved on the Press-Schechter formalism by assuming spherical collapse of Gaussian fluctuations or linear fluctuations that are surrounded by equal volumes of compensating underdensity. Except in the limit $`\beta 0`$, the PS formula overestimates the collapsed mass by factor $`(1.70)\beta ^{0.085}`$, or about 40% near $`\mathrm{\Omega }_{m0}=1/3`$. For simplicitly, this paper adheres to the PS formula with $`R_G=1`$ Mpc. In a forthcoming paper, we will use the improved MSW formula for both $`R_G=1,2`$ Mpc.) The variance of the mass power spectrum depends on the cosmological model ($`\mathrm{\Omega }_{m0}`$) and on the relevant co-moving galactic size scale $`R_G`$, but is insensitive to $`w_Q`$, for $`w_Q<1/3`$ . For the QCDM model we consider, $`\sigma _i^2(\mathrm{\Omega }_{m0},R_G)`$ is therefore the same as that already calculated for LCDM, for a scale-invariant mass spectrum smoothed with a top-hat window function. For the observed ratio $`_0=2,\mathrm{\Omega }_{m0}=1/3`$, at recombination $`1000\sigma =3.5,2.4`$, for comoving galactic size scale $`R_G=1,2`$ Mpc. The numerical factor in $`\delta _{i,c}`$ is $`9/5(4)^{1/3}=1.1339`$ for both $`w_Q=1`$ and $`w_Q=1/2`$, so that $`\delta _{i,c}=1.1339x/(1+z_i)`$. Thus, the collapsed mass parameter $`\sqrt{\beta }=0.80x/\sigma _i(R_G,u_0)`$, depends explicitly on $`u_0`$ through $`x=u_0,u_0^2`$ for LCDM, QCDM respectively. It also depends implicitly on $`u_0`$ through $`\sigma _i`$. Nevertheless, in going from LCDM the argument of $`f_{c,\mathrm{}}`$ scales simply as $`\sqrt{\beta }_{QCDM}=\sqrt{\beta }_{LCDM}u_0`$. Both asymptotic mass fractions are practically unity for large $`\mathrm{\Omega }_{m0}`$, but fall off with increasing ratio $`_0>1`$. For any $`_0>1`$, QCDM always leads to a smaller asymptotic mass fraction than LCDM. For ratio $`_0<1`$, $`f_{c,\mathrm{}}`$ changes slowly and the differences between QCDM and LCDM are not large. At the observed ratio $`R_0=2`$, the Press-Schechter asymptotic mass fractions are $`0.696,0.623`$ for LCDM, QCDM respectively. ### 3.2 Asymptotic Collapsed Mass Fraction Distribution Function As function of the ratio $`\mathrm{\Omega }_{m0}`$, the asymptotic mass fraction defines a distribution function $$f_{c,\mathrm{}}=d𝒫/d_0.$$ In Figure 2, instead of $`f_{c,\mathrm{}}`$ we plot the logarithmic distribution function in the ratio $`R_0`$ $$F(\mathrm{\Omega }_{m0})=_0f_{c,\mathrm{}}=d𝒫/d\mathrm{log}_0,$$ for LCDM and for QCDM and galactic size scale 1 Mpc. (Even for LCDM, this differs by a factor $`\sigma _i^3(\mathrm{\Omega }_{m0})`$ from the logarithmic distribution in $`\beta `$, $`d𝒫/d\mathrm{log}(\beta ^{3/2})`$ that is plotted by MSW and GLV.) $`F(\mathrm{\Omega }_{m0})`$ may be thought of as the ratio $`R_0`$ weighted by the number density of galaxies $`f_{c,\mathrm{}}`$. The figure shows broad peaks in the logarithmic distributions in $`\mathrm{\Omega }_{m0}`$ at ($`(\mathrm{\Omega }_{m0},F)=(0.23,1.27)`$ for LCDM and at $`(0.32,1.78)`$ for QCDM. At the observed $`\mathrm{\Omega }_{m0}=1/3`$, shown by circles (O), the asymptotic mass fraction logarithmic distribution in $`R_0`$ falls at 97% of the QCDM peak and at 78% of the LCDM peak. ## 4 $`\mathrm{\Omega }_Q\mathrm{\Omega }_m`$ is Quite Likely for Our Universe It is not surprising that our universe, containing at least one habitable galaxy, has $`_0=𝒪(1)`$. What is impressive is that our observed low-density universe, is almost exactly that which will maximize the number of habitable galaxies. Our existence does not explain $`\mathrm{\Omega }_{m0}`$, but the observed value makes our existence (and that of other evolved galaxies) most likely. What epistemological inference should we draw from this remarkable coincidence between our observed universe and the possible asymptotic mass fractions in either LCDM or QCDM universes? What should we infer statistically about any fundamental theory determining the parameters of our universe? An anthropic interpretation has already been given to “explain” a non-vanishing cosmological constant, in an assumed universe of subuniverses with all possible values for the vacuum energy $`\mathrm{\Omega }_\mathrm{\Lambda }=1\mathrm{\Omega }_{m0}`$. In each of these subuniverses, the probability for habitable galaxies to have emerged before the present epoch, is a function of $`\mathrm{\Omega }_\mathrm{\Lambda }`$: $$𝒫rob(\mathrm{\Omega }_{m0})(\text{prior distribution in}\mathrm{\Omega }_{m0})\times F(\mathrm{\Omega }_{m0}).$$ As always, the overall probability depends on the assumed prior. MSW, assuming nothing about initial conditions, take a prior flat in $`\mathrm{\Omega }_{m0}`$. GLV argue that the prior should be determined by a theory of initial conditions and is not flat for most theories. We prefer not to assume a distribution of real subuniverses, but to inversely apply Bayes’ Theorem to our own universe. In the absence of any physical explanation of the smooth energy, or until one is found, the partial information that intelligent astronomers exist tells us the observed smooth energy is just about what would be expected from equal a priori probabilities for $`\mathrm{\Omega }_{m0}`$ in the very early universe. That our universe is realized at or near the maximum in the asymptotic mass distribution function confirms that the prior is flat or peaked at $`\mathrm{\Omega }_{m0}1/3`$. Any phenomenologically viable fundamental theory must ultimately produce this value or be indifferent to the cosmological parameters. This research benefited from useful discussions with H. Martel and with M. Roos.
warning/0003/cond-mat0003502.html
ar5iv
text
# Two Kinds of the Coexistent States in One-Dimensional Quarter -Filled Systems under Magnetic Fields ## I Introduction Organic conductors such as (TMTSF)<sub>2</sub>$`X`$ and (TMTTF)<sub>2</sub>$`X`$ ($`X`$=ClO<sub>4</sub>, PF<sub>6</sub>, AsF<sub>6</sub>, ReO<sub>4</sub>, Br, SCN, etc.) are the quasi-one dimensional quarter-filled systems and exhibit various kinds of ground states, for example, spin-Peierls, spin density wave (SDW) and superconductivity. In (TMTSF)<sub>2</sub>PF<sub>6</sub>, the incommensurate SDW occurs at $`T=12`$ K, where the wave vector is determined by NMR experiments as (0.5, 0.24, 0.06). Pouget and Ravy observed the coexistence of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-charge density wave (CDW) by the X-ray scattering measurement, where $`k_\mathrm{F}=\pi /4a`$ is Fermi wave number and $`a`$ is the lattice constant. The critical temperature ($`T_{\mathrm{CDW}}`$) at which $`2k_\mathrm{F}`$-CDW is observed is the same temperature as $`T_{\mathrm{SDW}}`$. On the other hand, it is found that SDW in (TMTTF)<sub>2</sub>$`X`$ ($`X`$=Br and SCN) has the commensurate wave vector, (0.5, 0.25,0), as observed in the measurements of <sup>13</sup>C-NMR and <sup>1</sup>H-NMR. From the angle depenence of satelite peak positions of <sup>1</sup>H-NMR, the alignment of the spin moment along the conductive axis (a-axis) is obtained to be ($`,0,,0`$), where the arrow means the spin moment. In (TMTTF)<sub>2</sub>Br, $`4k_\mathrm{F}`$-CDW accompanied by $`2k_\mathrm{F}`$-SDW is found in X-ray scattering measurments, where $`T_{\mathrm{CDW}}100`$ K and $`T_{\mathrm{SDW}}13`$ K. In the one-dimensional quarter-filled systems, the CDW-SDW coexistent state is shown to be caused by the interplay between the on-site Coulomb interaction ($`U`$) and the inter-site Coulomb interaction ($`V`$). The inter-site Coulomb interaction plays important role for the charge order. Mila has estimated $`U/t5`$ and $`V/t2`$, where $`t`$ is a transfer integral. Seo and Fukuyama showed that the ground state becomes the coexistent state of $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW in the one-dimensional extended Hubbard model with $`V`$. It is found by Kobayashi et al. and Mazumdar et al. that the coexistent state of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-CDW is stabilized when the next nearest neighbor Coulomb interaction ($`V_2`$) and the dimerization of the energy band are considered. Thus, two kinds of the coexistent states ($`2k_\mathrm{F}`$-SDW-$`2k_\mathrm{F}`$-CDW and $`2k_\mathrm{F}`$-SDW-$`4k_\mathrm{F}`$-CDW), which are observed by the X-ray scattering mesauerments in (TMTSF)<sub>2</sub>PF<sub>6</sub> and (TMTTF)<sub>2</sub>Br, can be explained by Seo and Fukuyama and Kobayashi et al. and Mazumdar et al., respectively. Recently, we have studied the effects of the magnetic field ($`H`$) on the coexistent state of $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW for both cases of strong and weak coupling of the correlation between electrons, where the one-dimensional quarter-filled extended Hubbard model with $`V`$ is used. It is found that although the spin order is suppressed at high fields, the charge order still remains in the strongly coupling systems ($`U/t5`$). When the coupling is small ($`U/t1.5`$), both orderings of $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW disappear at the critical magnetic field, which is the same as the Pauli paramagnetic field in the spin-singlet superconductivity. In this paper, we examine how the coexistent state of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-CDW and the one of $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW are affected by magnetic fields and temperatures. We use the one-dimensional extened Hubbard model with $`V`$ and $`V_2`$, where the dimerization is neglected for simplicity, because the coexistent states are stabilized without the dimerization. We use parameters, $`U/t=5.0`$, which is indicated in (TMTTF)<sub>2</sub>$`X`$ ((TMTSF)<sub>2</sub>$`X`$). In most cases we set $`V/U1`$ and $`V_2<V`$. We study the case when the anisotorpy of the spin is strong, because it is found that the easy axis of the spin in the quasi-one-dimensional organic conductors is $`b`$-axis. In order to clarify the ground state under fields at finite temperatures, we show the $`V`$-$`H`$, $`V_2`$-$`H`$, $`V`$-$`T`$ and $`V_2`$-$`T`$ phase diagrams, which enables to discuss the effect of pressure and the ordering temperatures of the SDW and CDW. It is expected that the Hubbard model with large $`U/t5`$ is qualitatively understood by the Ising model with the perpendicular field, as we will show below. The $`H`$-dependence of the antiferromagnetic state in one-dimensional Ising model has been studied in the mean field approximation, where the componet of the spin along the $`c`$-axis is considered. When the magnetic fields is applied perpendicular to the spin moment, $`S_z(H,j)/S_z(0,j)=\sqrt{1(H/H_x^0)^2},`$ (1) where $`S_z(H,j)`$ is the amplitude of the spin moment at $`j`$ site at $`H=0`$ and $`H_x^0`$ is the critical field at which the ordering of the antiferromagnetic state disappear ($`H_x^0J`$, where $`J`$ is the coupling constant). ## II Formulation We study the one-dimensional extended Hubbard model, $`\widehat{}`$ $`=`$ $`\widehat{𝒦}+\widehat{𝒰}+\widehat{𝒱}+\widehat{𝒱}_2,`$ (2) $`\widehat{𝒦}`$ $`=`$ $`t{\displaystyle \underset{i,\sigma }{}}(c_{i,\sigma }^{}c_{i+1,\sigma }+h.c.)`$ (4) $`+{\displaystyle \frac{\mu _\mathrm{B}g}{2}}{\displaystyle \underset{i,\sigma ,\sigma ^{}}{}}c_{i,\sigma }^{}(H_x)_{\sigma \sigma ^{}}c_{i,\sigma ^{}},`$ $`\widehat{𝒰}`$ $`=`$ $`U{\displaystyle \underset{i}{}}n_{i,}n_{i,},`$ (5) $`\widehat{𝒱}`$ $`=`$ $`V{\displaystyle \underset{i,\sigma ,\sigma ^{}}{}}n_{i,\sigma }n_{i+1,\sigma ^{}},`$ (6) $`\widehat{𝒱}_2`$ $`=`$ $`V_2{\displaystyle \underset{i,\sigma ,\sigma ^{}}{}}n_{i,\sigma }n_{i+2,\sigma ^{}},`$ (7) where $`\mu _\mathrm{B}=e\mathrm{}/2m_0c`$ is the Bohr magneton and $`c_{i,\sigma }^{}`$ is the creation operator of $`\sigma `$ spin electron at $`i`$ site, $`n_{i,\sigma }`$ is the number operator, $`g=2`$, $`i=1,\mathrm{},N_\mathrm{S}`$, $`N_\mathrm{S}`$ is the number of the total sites and $`\sigma =`$ and $``$. The notation in this paper follows Seo and Fukuyama. The magnetic field is applied to the x-axis and $`(H_x)_{\sigma \sigma ^{}}`$ $`=H(\widehat{\sigma }_x)_{\sigma \sigma ^{}}`$, where $`H`$ is the strength of the magnetic field and $`\widehat{\sigma }_x`$ is Pauli spin matrix. We consider the quarter-filled case. The interaction terms, $`\widehat{𝒰}`$, $`\widehat{𝒱}`$ and $`\widehat{𝒱}_2`$ are treated in the mean field approximation as $`\widehat{𝒰}^\mathrm{M}`$ $`=`$ $`{\displaystyle \underset{k_x}{}}{\displaystyle \underset{Q}{}}\{\rho _{}(Q,T)C^{}(k_x,)C(k_xQ,)`$ (8) $`+`$ $`\rho _{}^{}(Q,T)C^{}(k_xQ,)C(k_x,)\}`$ (9) $`\widehat{𝒱}^\mathrm{M}`$ $`=`$ $`({\displaystyle \frac{V}{U}}){\displaystyle \underset{k_x,\sigma ,\sigma ^{}}{}}{\displaystyle \underset{Q}{}}e^{iQa}\{\rho _{\sigma \sigma }(Q,T)C^{}(k_x,\sigma ^{})C(k_xQ,\sigma ^{})`$ (10) $`+`$ $`\rho _{\sigma ^{}\sigma ^{}}^{}(Q,T)C^{}(k_x,\sigma )C(k_xQ,\sigma )\}`$ (11) $`\widehat{𝒱}_{2}^{}{}_{}{}^{\mathrm{M}}`$ $`=`$ $`({\displaystyle \frac{V_2}{U}}){\displaystyle \underset{k_x,\sigma ,\sigma ^{}}{}}{\displaystyle \underset{Q}{}}e^{2iQa}\{\rho _{\sigma \sigma }(Q,T)C^{}(k_x,\sigma ^{})C(k_xQ,\sigma ^{})`$ (12) $`+`$ $`\rho _{\sigma ^{}\sigma ^{}}^{}(Q,T)C^{}(k_x,\sigma )C(k_xQ,\sigma )\}`$ (13) where $`I=U/N_\mathrm{S}`$. The self-consistent equation for the order parameter $`\rho _{\sigma \sigma }(Q,T)`$ at finite temperature, $`T`$, is given by $`\rho _{\sigma \sigma }(Q,T)`$ $`=`$ $`I{\displaystyle \underset{k_x}{}}<C^{}(k_x,\sigma )C(k_xQ,\sigma )>.`$ (14) In eqs. (7)-(10), we take the order parameters, $`\rho _{\sigma \sigma }(Q,T)`$, only between electrons with same spins. We do not consider the case of the mean field, $`\rho _{\sigma \overline{\sigma }}(Q,T)=I_{k_x}<C^{}(k_x,\sigma )C(k_xQ,\overline{\sigma })>`$ with $`\sigma \overline{\sigma }`$. We neglect the effect of the spin tilting to the $`xy`$ plane by setting $`\rho _{\sigma \overline{\sigma }}(Q,T)=0`$ in this mean field approximation. This simplification corresponds to the assumption that the rotational symmetry in the spin-space is broken and that the $`z`$-axis is the easy axis although the spin in this Hubbard model is isotropic. The magnetic field is applied perpendicular to the easy axis when $`H0`$. In the Fulde-Ferrel-Larkin-Ovchinnikov state in the superconducutivity, the total moment of the Cooper pair is changed by the magnetic field. Similar situation may occur in SDW and CDW, i.e., the wave vector $`Q`$ may be changed by the magnetic field. However, such state will be stabilized only in the low temperature and strong magnetic field region and will not change the essential feature studied in this paper. Therefore, we take the possible wave vectors of the order parameters as $`Q=Q_0,2Q_0,3Q_0`$ and $`4Q_0`$ (equivalent to 0), where $`Q_0=2k_\mathrm{F}`$. We evaluate the Helmholtz free energy by using the eigenvalue $`ϵ_j`$ and the unitary matrix $`U_{𝐤\sigma ,j}`$ obtained by diagonalizing the mean field Hamiltonian $`\widehat{𝒦}+\widehat{𝒰}^\mathrm{M}+\widehat{𝒱}^\mathrm{M}+\widehat{𝒱}_2^\mathrm{M}`$, where index $`j`$ includes the degree of the spin freedom. We determine the chemical potential $`\mu (\rho _{\sigma \sigma }(Q,T))`$ from $`N=1/4`$, where $`N`$ $`=`$ $`{\displaystyle \underset{k_x,\sigma }{}}<C^{}(k_x,\sigma )C(k_x,\sigma )>`$ (15) $`=`$ $`{\displaystyle \frac{1}{2N_\mathrm{s}}}{\displaystyle \underset{j=1}{\overset{2N_\mathrm{s}}{}}}\left[\mathrm{exp}\left({\displaystyle \frac{ϵ_j\mu }{T}}\right)+1\right]^1.`$ (16) The self-consistency condition (eq. (10)) is given by $`\rho _{\sigma \sigma }(Q,T)=I{\displaystyle \underset{k_x}{}}{\displaystyle \underset{j}{}}U_{(k_x,\sigma ),j}^{}U_{(k_xQ,\sigma ),j}f(ϵ_j),`$ (17) where $`f(ϵ_j)`$ is the Fermi function. We obtain the free energy per site $`F(\rho _{\sigma \sigma }(Q,T))`$ $`=`$ $`2\mu n{\displaystyle \frac{T}{N_\mathrm{s}}}{\displaystyle \underset{j=1}{\overset{2N_\mathrm{s}}{}}}\mathrm{log}\left\{\mathrm{exp}\left({\displaystyle \frac{\mu ϵ_j}{T}}\right)+1\right\}`$ (18) $``$ $`{\displaystyle \frac{1}{U}}{\displaystyle \underset{Q}{}}\rho _{}(Q,T)\rho _{}^{}(Q,T),`$ (19) $``$ $`{\displaystyle \frac{V}{U^2}}{\displaystyle \underset{Q,\sigma ,\sigma ^{}}{}}e^{iQa}\rho _{\sigma \sigma }(Q,T)\rho _{\sigma ^{}\sigma ^{}}^{}(Q,T)`$ (20) $``$ $`{\displaystyle \frac{V_2}{U^2}}{\displaystyle \underset{Q,\sigma ,\sigma ^{}}{}}e^{2iQa}\rho _{\sigma \sigma }(Q,T)\rho _{\sigma ^{}\sigma ^{}}^{}(Q,T),`$ (21) At $`T=0`$ it reduces to the ground state energy per site $`E(H,\rho _{\sigma \sigma }(Q,0))=`$ $`{\displaystyle \frac{1}{N_s}}{\displaystyle \underset{j=1}{\overset{N_s/2}{}}}ϵ_j{\displaystyle \frac{1}{U}}{\displaystyle \underset{Q}{}}\rho _{}(Q,0)\rho _{}^{}(Q,0),`$ (22) $``$ $`{\displaystyle \frac{V}{U^2}}{\displaystyle \underset{Q,\sigma ,\sigma ^{}}{}}e^{iQa}\rho _{\sigma \sigma }(Q,0)\rho _{\sigma ^{}\sigma ^{}}^{}(Q,0)`$ (23) $``$ $`{\displaystyle \frac{V_2}{U^2}}{\displaystyle \underset{Q,\sigma ,\sigma ^{}}{}}e^{2iQa}\rho _{\sigma \sigma }(Q,0)\rho _{\sigma ^{}\sigma ^{}}^{}(Q,0).`$ (24) The electron density at the $`j`$ site, $`n(j,T)`$, its deviation from the mean value (1/2), $`\delta (j,T)`$, and the spin moment at $`j`$ site, $`S_z(j,T)`$, are given by $`n(j,T)={\displaystyle \frac{1}{U}}{\displaystyle \underset{Q,\sigma }{}}\rho _{\sigma \sigma }(Q,T)e^{iQja}={\displaystyle \frac{1}{2}}+\delta (j,T),`$ (25) and $`S_z(j,T)={\displaystyle \frac{1}{2U}}{\displaystyle \underset{Q}{}}(\rho _{}(Q,T)\rho _{}(Q,T))e^{iQja}.`$ (26) When $`T=0`$, $`n(j)n(j,0)`$, $`S_z(j)S_z(j,0)`$ and $`\delta (j)\delta (j,0)`$. ## III Results and Discussions ### A 2k<sub>F</sub>-SDW and 4k<sub>F</sub>-CDW In this subsection, we take $`V_2=0`$ and $`T=0`$. It is known that the coexistent state of 2$`k_\mathrm{F}`$-SDW and 4$`k_\mathrm{F}`$-CDW are induced by $`U`$ and $`V`$.. Figs. 1 and 2 are $`S_z`$ and $`\delta `$ as a function of $`V/t`$ at $`U/t=5.0`$ and $`H=0`$. For $`0V0.39`$, the antiferromagnetic order (($``$,$``$,$``$,$``$), i.e., $`S_z(1)=S_z(2)=S_z(3)=S_z(4)`$) is stabilized and there is no charge order ($`\delta (1)=\delta (2)=\delta (3)=\delta (4)=0`$). The spin order of ($``$,$``$,$``$,$``$) has the wave vector of $`2k_\mathrm{F}`$. For $`V/t>0.39`$, the spin order becomes ($``$,0,$``$,0) ($`S_z(1)=S_z(3)`$, $`S_z(2)=S_z(4)=0`$) and the charge order ($`\delta `$,$`\delta `$,$`\delta `$,$`\delta `$) coexists, where $`\delta (1)=\delta (3)=\delta `$ and $`\delta (2)=\delta (4)=\delta `$. These orders, ($``$,0,$``$,0) and ($`\delta `$,$`\delta `$,$`\delta `$,$`\delta `$), mean $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW, respectively. These results were obtained by Seo and Fukuyama. We show $`S_z`$ and $`\delta `$ for $`U/t=5.0`$ and $`V/t=05.0`$ as a function of perpendicular field ($`h_x\mu _\mathrm{B}H/t`$) in Figs. 3 and 4. The antiferromagnetic state is gradually suppressed up to the critical field ($`h_x^c`$=1.3, 1.3, 1.5, 1.8 and 2.4 for $`V/t`$=0, 1.0, 2.0, 2.5 and 5.0, respectively), as shown in Fig. 3, where $`S_z(1)=S_z(3)`$ and $`S_z(2)=S_z(4)=0`$. Upon increasing $`h_x`$ the charge order decreases and becomes zero when $`V/t=1.0`$, and the charge order becomes constant for $`h_x>h_x^c`$ when $`V/t1.5`$, as shown in Fig. 4, where $`\delta (1)=\delta (3)=\delta `$ and $`\delta (2)=\delta (4)=\delta `$. The $`h_x`$-dependence of the amplitude of $`S_z`$ is almost the same as eq. (1) when we set $`H_x^0`$ as $`H_x^0=th_x^c/\mu _\mathrm{B}`$, which is shown by solid lines in Fig. 3. From Figs. 3 and 4, we make the $`V`$-$`h_x`$ phase diagram, as shown in Fig. 5, where (0,0,0,0) means that the spin and charge orders do not exist. The dotted lines show the second order transition lines and the solid line show the first order transition lines. In the region of large $`h_x`$ and $`V`$, the state of $`4k_\mathrm{F}`$-CDW, ($`\delta `$,$`\delta `$,$`\delta `$,$`\delta `$), is stabilized. The $`h_x`$ -dependence of the spin order can be understood by the mean field solutions of the Ising model mentioned in the introduction. ### B 2k<sub>F</sub>-SDW and 2k<sub>F</sub>-CDW Here, $`V_2`$ is introduced, which favors the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW. We search the most stable solution at $`U/t=5.0`$, $`V/t=2.0`$, $`T=0`$ and $`H=0`$ by changing $`V_2`$. These $`V_2`$-dependences of the spin density and charge density are shown in Figs. 6 and 7. For $`0V_2/t1.4`$, the coexistent state of 2$`k_\mathrm{F}`$-SDW and 4$`k_\mathrm{F}`$-CDW is stable, while the state changes to the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW ((,$``$,$``$,), i.e., $`S_z(1)=S_z(4)<S_z(2)=S_z(3)`$) and (($`\delta `$,$`\delta `$,$`\delta `$,$`\delta `$), $`\delta (1)=\delta (4)=\delta `$ and $`\delta (2)=\delta (3)=\delta `$) for $`V_2/t>1.4`$. This transition is a first order transition. The coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW obtained by Kobayashi et al. is ($``$,$``$,$``$,$``$) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$), which is different from the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW ((,$``$,$``$,) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$)). Since they limited the freedom of the order parameters, they could not find this coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW ((,$``$,$``$,) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$)). They also indicated that the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW (($``$,$``$,$``$,$``$) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$)) is suppressed when the dimerization is not included. In fact, we could not find the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW of ($``$,$``$,$``$,$``$) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$) in the parameter region, ($`U/t=5.0`$, $`V/t=2.0`$ and $`V_2/t<2.0`$). However, as we show here, the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW ((,$``$,$``$,) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$)) is stabilized even in the absence of the dimerization. After we finished this study, we know the very recent study at $`T=0`$ and $`T0`$ in the absence of the magnetic field by Tomio and Suzumura. They calculate the mean field solution using the extended Hubbard model with $`V`$ and $`V_2`$. They show that the state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW ((,$``$,$``$,) and ($`\delta `$, $`\delta `$, $`\delta `$, $`\delta `$)) coexists even if the dimerization is not included, too. We analyze the coexistent state of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-CDW under magnetic fields at $`0V_2/t2.0`$. In the case of the perpendicular field ($`h_x`$), the state of (,$``$,$``$,) and ($`\delta `$,$`\delta `$,$`\delta `$,$`\delta `$) becomes the one of ($``$,$``$,$``$,$``$) and (0,0,0,0) ($`\delta (1)=\delta (2)=\delta (3)=\delta (4)=0`$) at $`0.75<h_x<1.25`$, which is shown in Figs. 8 and 9. We can see that the coexistent state of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-CDW changes to the state of $`2k_\mathrm{F}`$-SDW. This transitions is the second order phase transition, since $`|S_z(1)|`$ and $`|S_z(3)|`$ ($`|S_z(2)|`$ and $`|S_z(4)|`$) increase (decrease) gradually, and (,$``$,$``$,) changes to ($``$,$``$,$``$,$``$) smoothly. When $`h_x>1.25`$, the $`2k_\mathrm{F}`$-SDW state disappears. For $`0V_2/t1.4`$, as $`h_x`$ increases, the spin density in the coexistent state of $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW decrease and becomes zero at $`h_x^c=1.3,1.3,1.3`$ and 1.5 for $`V_2/t=1.4,1.0,0.5`$ and 0, as shown in Fig. 10, where $`S_z(1)=S_z(3)`$ and $`S_z(2)=S_z(4)=0`$. When we set $`H_x^0`$ as $`H_x^0=th_x^c/\mu _\mathrm{B}`$, the $`h_x`$-dependence of $`S_z`$ is in agreement with eq. (1), which is shown by solid line in Fig. 10. The order of $`4k_\mathrm{F}`$-CDW becomes nonzero at the higher field for $`V_2/t<0.5`$, as shown in Fig. 11, where $`\delta (1)=\delta (3)=\delta `$ and $`\delta (2)=\delta (4)=\delta `$. We show the $`V_2`$-$`h_x`$ phase in Fig. 12, where the solid line is for the first order transition and the dotted lines are for the second order transitions. ### C Finite Temperature We show $`S_z`$ and $`\delta `$ as a function of $`T`$ at $`H=0`$. At $`U/t=5.0`$ and $`V_2/t=0`$, $`S_z`$ and $`\delta `$ as a function of $`T`$ for various values of $`V`$ are shown in Figs 13 and 14, where $`S_z(1,T)=S_z(2,T)=S_z(3,T)=S_z(4,T)`$ for $`V/t=0`$ and $`S_z(1,T)=S_z(3,T),S_z(2,T)=S_z(4,T)=0`$, $`\delta (1,T)=\delta (3,T)=\delta (T)`$ and $`\delta (2,T)=\delta (4,T)=\delta (T)`$ for $`V/t=0.55.0`$. As $`V`$ increases, the critical temperatures ($`T_{\mathrm{SDW}}`$ and $`T_{\mathrm{CDW}}`$) at which $`S_z`$ and $`\delta `$ become zero increase. The $`T`$-dependences of $`n`$ for $`V/t=3.0`$, 4.0 and 5.0 have a dip at $`T_{\mathrm{SDW}}`$ for $`V/t=3.0`$, 4.0 and 5.0, which can be seen in Fig. 14. We show the $`V`$-$`T`$ phase diagram in Fig. 15, where the solid and dotted lines are for the first and the second order transitions, respectively. Note that $`T_{\mathrm{CDW}}>T_{\mathrm{SDW}}`$ for $`V/t>2.0`$, that is, the charge order remains even if the spin order disappears under high temperatures. As we have shown above, the coexistent state of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-CDW ((,$``$,$``$,) and ($`\delta `$,$`\delta `$,$`\delta `$,$`\delta `$)) are stabilized for $`V_2/t>1.4`$ at $`T/t=0`$ (see Figs. 6 and 7). At $`T/t0`$, for example, for $`V_2/t=1.6`$ as $`T`$ increases, the magnitudes of $`S_z`$ and $`n`$ decreases and the second order transition to the state of $`2k_\mathrm{F}`$-SDW (($``$,$``$,$``$,$``$)) occurs, as shown in Figs. 16 and 17. Namely, $`T_{\mathrm{CDW}}T_{\mathrm{SDW}}`$ in the coexistent state of $`2k_\mathrm{F}`$-SDW and $`2k_\mathrm{F}`$-CDW. For $`V_2/t<1.4`$, $`T_{\mathrm{CDW}}`$ in the coexistent state of $`2k_\mathrm{F}`$-SDW and $`4k_\mathrm{F}`$-CDW is the same as $`T_{\mathrm{SDW}}`$, as shown in Figs. 18 and 19, where $`S_z(1,T)=S_z(3,T),S_z(2,T)=S_z(4,T)=0`$, $`\delta (1,T)=\delta (3,T)=\delta (T)`$ and $`\delta (2,T)=\delta (4,T)=\delta (T)`$. We show the $`V_2`$-$`T`$ phase diagram in Fig. 20, where the solid and dotted lines are for the first and the second order transition lines, respectively. Our phase diagrams ($`V`$-$`T`$ and $`V_2`$-$`T`$) are in good agreement with those obtained by Tomio and Suzumura. ## IV Conclusions We have shown the different dependences of the magnetic field and temperatures on the 2$`k_\mathrm{F}`$-SDW-2$`k_\mathrm{F}`$-CDW coexistent state and 2$`k_\mathrm{F}`$-SDW-4$`k_\mathrm{F}`$-CDW coexisting state. When $`V_2`$ is absent or small, the coexistent state of 2$`k_\mathrm{F}`$-SDW and 4$`k_\mathrm{F}`$-CDW is stable. Then the charge order survives at high fields in the case of the perpendicular magnetic field to the easy axis when $`V/t>2.0`$ . This situation may be realized in (TMTTF)<sub>2</sub>$`X`$, where 4$`k_\mathrm{F}`$-CDW is expected to be observed in X-ray scattering measurements under high fields. On the other hand, if $`V_2`$ is large enough, the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW is stabilized. When the magnetic field is applied perpendicular along the easy axis, the coexistent state of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW changes to 2$`k_\mathrm{F}`$-SDW state. This transition is the second order phase transition, and (,$``$,$``$,) becomes ($``$,$``$,$``$,$``$). This transition may be observed in the angle dependence of satelite peak positions of NMR in (TMTSF)<sub>2</sub>PF<sub>6</sub>. The critial temperature of the charge order ($`T_{\mathrm{CDW}}`$) is higher than that of the spin order ($`T_{\mathrm{SDW}}`$) in the coexistent phase of 2$`k_\mathrm{F}`$-SDW and 4$`k_\mathrm{F}`$-CDW when $`V/t>2.0`$. This is in good agreement with $`T_{\mathrm{SDW}}13`$ K and $`T_{\mathrm{CDW}}100`$ K observed in X-ray scattering in (TMTTF)<sub>2</sub>Br. Since $`T_{\mathrm{CDW}}`$ is convergent to $`T_{\mathrm{SDW}}`$ for $`V/t<2.0`$, we expect that the critical temperatures of 2$`k_\mathrm{F}`$-SDW and 4$`k_\mathrm{F}`$-CDW will become the same when the pressure is applied to (TMTTF)<sub>2</sub>Br. In the coexistent phase of 2$`k_\mathrm{F}`$-SDW and 2$`k_\mathrm{F}`$-CDW, we find that $`T_{\mathrm{CDW}}T_{\mathrm{SDW}}`$, which is consistent with the X-ray scattering measurement in (TMTSF)<sub>2</sub>PF<sub>6</sub>, where $`T_{\mathrm{CDW}}T_{\mathrm{SDW}}`$. ## V Acknowledgment The authors would like to thank T. Sakai, K. Machida, Y. Nogami, Y. Suzumura and Y. Tomio for valuable discussions.
warning/0003/astro-ph0003427.html
ar5iv
text
# HST spectroscopy of the double QSO HS 1216+5032 ABBased on observations with the NASA/ESA Hubble Space Telescope, obtained at the STScI, which is operated by AURA, Inc., under NASA contract NAS5–26 555. ## 1 Introduction HS 1216+5032 was discovered in the course of the Hamburg Quasar Survey (Hagen et al. Hagen1 (1995)). It was later confirmed as a double QSO at $`z_e=1.45`$ by Hagen et al. (Hagen (1996)) through direct images and low-resolution (FWHM $`20`$ Å) spectroscopy in the optical range. The $`B`$ magnitudes of the bright and faint QSO images (hereafter “A” and “B”, respectively) are $`B_\mathrm{A}=17.2`$ and $`B_\mathrm{B}=19.0`$. The separation angle between A and B is $`\theta =9\stackrel{}{.}1`$, and the emission redshifts are $`z_e(\mathrm{A})=1.455`$ and $`z_e(\mathrm{B})=1.451`$. From the available optical data of HS 1216+5032 AB there appears to be some evidence favoring its physical-pair nature instead of a gravitational lens origin (this subject will be discussed in §6.1), so the projected distance<sup>1</sup><sup>1</sup>1Throughout this paper we assume an $`(\mathrm{\Omega }_0,\mathrm{\Lambda }_0)=(1,0)`$ cosmology and define $`h_{50}H_0/(50`$ km s<sup>-1</sup>Mpc$`{}_{}{}^{1})`$. between LOSs as a function of redshift is known ($`78h_{50}^1`$ kpc at $`z_e`$). This paper presents ultraviolet (UV) spectra of QSO A and B taken with the Faint Object Spectrograph (FOS) on-board the Hubble Space Telescope (HST). The prime goal of these observations was originally to investigate the Ly$`\alpha `$ forest in the little explored redshift range between $`z=0.8`$ and $`z=1.4`$. In particular, the projected separation between the LOSs to HS 1216+5032 A and B is convenient because it samples well the expected Ly$`\alpha `$ cloud sizes of hundreds of kpc. For the redshift range accessible from the ground, a pioneering work on cloud sizes was made by Smette et al. (Smette1 (1992)) using the gravitationally lensed double QSO UM 673 ($`\theta =2\stackrel{}{.}2`$) with known lens geometry. All observed Ly$`\alpha `$ lines were observed to be common to both spectra and a statistical lower limit of $`D>12`$ $`h_{50}^1`$ kpc could be set for the transverse sizes (spherical clouds, no evolution). A similar result (Smette et al. Smette (1995)) was found for HE 1104$``$1805 AB ($`\theta =3\mathrm{}`$) but here the position of the lens $`z_\mathrm{l}`$ was unknown. A transverse size of $`D>60h_{50}^1`$ kpc was estimated for $`z_\mathrm{l}1`$. Surprisingly, QSO pairs at very large projected distances seem to still show Ly$`\alpha `$ lines common to both spectra. In the spectra of LB 9605 and LB 9612 ($`\theta =1\stackrel{}{.}65`$), Dinshaw et al. (Dinshaw1 (1998)) find 5 such lines within $`400`$ km s<sup>-1</sup> and derive a most probable diameter of $`1520`$ $`h_{50}^1`$ kpc at $`1<z<1.7`$. If the lines arise in different absorbers, however, an upper limit of $`1140`$ $`h_{50}^1`$ kpc is derived. Further studies in the optical range have been carried out by Dinshaw et al. (Dinshaw (1994); $`\theta =9\stackrel{}{.}5`$, $`D>160`$ $`h_{50}^1`$ kpc at $`1.7<z<2.1`$) and Crotts et al. (Crotts (1994); same results). At low redshift ($`z1.5`$), Ly$`\alpha `$ clouds show a very flat redshift distribution: only $`16\pm 2`$ Ly$`\alpha `$ lines with $`W_0>0.32`$ Å are expected in the wavelength interval covered by the FOS on the HST (Weymann et al. Weymann1 (1998)). Therefore, the size-estimate uncertainties for low-redshift Ly$`\alpha `$ absorbers are large. Studies using FOS spectra of QSO pairs have been made by Dinshaw et al. (Dinshaw2 (1997); $`\theta =1\stackrel{}{.}29`$, $`D>600`$ $`h_{50}^1`$ kpc at $`0.5<z<0.9`$), and more recently by Petitjean et al. (Petitjean (1998); $`\theta =36\mathrm{}`$, $`D500`$ $`h_{50}^1`$ kpc at $`0.8<z<1.4`$). All the aforementioned limits on Ly$`\alpha `$ cloud sizes result from rather crude models which assume non-evolving and uniform-sized absorbers. Such simple models might be unrealistic, as suggested for example by a trend of larger size estimates with larger LOS separation (Fang et al. Fang (1996)) or by the claim (Fang et al. Fang (1996), Dinshaw et al. Dinshaw1 (1998)) that the cloud size increased with cosmic time. However, we note that none of the above claims is confirmed by D’Odorico et al. (D'Odorico (1998)), using a larger database. Clearly, the issues of geometry and size of Ly$`\alpha `$ absorbers remain open due partly to a lack of size estimates at low redshift. In this paper we use a likelihood analysis (§ 6) that attempts to constrain the size of Ly$`\alpha `$ clouds in front of HS 1216+5032 A and B based on simple considerations on shape and evolution. We model the absorbers either as non-evolving spheres or as filaments. The technique takes advantage of the information provided by line pairs observed in both spectra at the same redshift, hereafter referred to as “coincidences”, and Ly$`\alpha `$ lines observed only in one spectrum, referred to as “anti-coincidences”. A likelihood function is then constructed using an analytic expression for the probability of getting the observed number of coincidences and anti-coincidences in each of the two geometries. In general, modelling the geometry of Ly$`\alpha `$ absorbers at low redshift on the basis of line counting is made difficult because the line samples are intrinsically small, implying a poor statistics. In addition, HS 1216+5032 B shows broad absorption lines (BALs) caused by several ions with transitions in the UV, so the effective redshift path for the detection of Ly$`\alpha `$ lines is even shorter. Intervening metal absorption lines of interest are presented in § 4. The BAL systems are described in § 5. The conclusions are outlined in § 7. ## 2 Data reduction ### 2.1 Observations and wavelength calibration UV spectra of HS 1216+5032 A and B were obtained on 1996 November 6 with the FOS on board the HST. Target acquisition and spectroscopy were done using Grating G270H with the red detector and the $`3.{}_{}{}^{\prime \prime }7\times 3.{}_{}{}^{\prime \prime }7`$ aperture. This configuration yields a spectral resolution of FWHM $`=2`$ Å and a wavelength coverage from 2222 Å to 3277 Å (Schneider et al. Schneider (1993)). Total integration times were 1980 and 10400 s for QSO components A and B, respectively. The spectrum of image B resulted from the variance-weighted addition of four exposures. The signal-to-noise ratios at Ly$`\alpha `$ emission are S/N $`=40`$ (A) and $`25`$ per $`0.5`$ Å pixel, falling down to 15 in the blue part of the A spectrum, and to near one at the BAL troughs in B. The flux-calibrated spectra and their associated 1 $`\sigma `$ errors are shown in Figure 1. The dotted lines represent the continua (§ 2.3) and the tick-marks indicate the position of $`3\sigma `$ absorption lines (§ 3). Since we are interested in comparing absorption systems along both LOSs, we must check for possible small misalignments of the zero-point wavelength scale between both spectra. Such differences may arise when the targets have not been properly centered in the aperture of the FOS, and/or when – unlike here – the science and wavelength calibration exposures have not been performed consecutively. Galactic absorption lines common to both spectra were then required to appear at the same wavelength, under the assumption that both LOSs cross the same cloud (which should be true for the small separation between LOSs at $`z0`$). There are two Galactic absorption lines, Fe ii $`\lambda 2600`$ and Mg ii $`\lambda 2796`$, common to both spectra and apparently not contaminated with intergalactic lines. For these lines $`\mathrm{\Delta }\lambda `$(Fe ii $`\lambda 2600`$)$`=9`$ km s<sup>-1</sup> and $`\mathrm{\Delta }\lambda `$(Mg ii $`\lambda 2796`$)$`=+4`$ km s<sup>-1</sup> are obtained, which lie within the fitting procedure errors ($`10`$ km s<sup>-1</sup>) and the FOS limiting accuracy ($`20`$ km s<sup>-1</sup>). These small differences imply that both spectra are well-calibrated; consequently, we made no corrections to the zero-point wavelength scale. ### 2.2 Emission redshifts To derive the emission redshift of HS 1216+5032 A the Ly$`\alpha `$ emission peak was fitted with a Gaussian in the region between 2959 and 3005 Å. Excluding the wavelength regions with absorption features in the blue wing of the line yields $`z_e(\mathrm{A})=1.4545\pm 0.0001`$, where the $`1\sigma `$ error comes from the wavelength uncertainty. For QSO component B, a similar analysis is extremely difficult because the Ly$`\alpha `$ emission line is severely distorted by the Ly$`\alpha `$ and N v BALs. Fitting instead the N v emission peak between 3016 and 3065 Å, and using the mean rest-frame wavelength of the doublet $`\lambda _0`$(N v)$`=1240.15`$ Å, yields $`z_e(\mathrm{B})=1.4509\pm 0.0006`$. In the following we adopt the nominal values $`z_e`$(A)$`=1.455`$ and $`z_e`$(B)$`=1.451`$. The redshifts of A and B are not the same within measurement errors, with QSO component B being blueshifted by $`400`$ km s<sup>-1</sup> relative to A. Different emission redshifts are expected in the physical pair hypothesis, but the measured difference between HS 1216+5032 A and B cannot rule out the possibility that the QSO may still be gravitationally lensed. This is because two different transitions, Ly$`\alpha `$ and N v, are being compared and blueshifts of a few hundreds km s<sup>-1</sup> between high and low-ionization emission lines are common (e.g., Hamann et al. Hamman (1997); Tytler & Fan Tytler (1992)). This point is further discussed in § 6.1. Other prominent emission lines in the spectrum of A are: O vi $`\lambda 1033`$ (blended with Ly$`\beta `$) and probably C iii $`\lambda 977`$ and S iv $`\lambda 1062`$. In the spectrum of B, the BAL troughs allow clear identification of only Ly$`\alpha `$ and N v $`\lambda 1240`$, but the “undulating” shape of the observed flux for $`\lambda <2600`$ Å suggests that S vi $`\lambda 939`$, C iii $`\lambda 977`$, N iii $`\lambda 990`$, O vi $`\lambda 1033`$, and S iv $`\lambda 1062`$ have broad absorption troughs on the blue side of the expected QSO rest wavelengths. ### 2.3 Continuum fitting Due to line blending in the Ly$`\alpha `$ forest, defining a quasar continuum shortward of Ly$`\alpha `$ emission is not trivial at our resolution. For B, this difficulty is increased by the BAL profiles. A power-law shaped continuum $`f\nu ^\alpha `$, although an acceptable representation in the optical range, is definitively not able to fit even the UV continuum spectrum of quasar image A, leaving residuals far larger than 3 $`\sigma `$ at apparently featureless spectral regions. We decided to fit the UV QSO continuum for each spectrum with cubic splines, using the following simple and semi-automated fit algorithm. First, cubic splines are fitted through a given number of points $`(x_i,y_i)`$ thought to represent the continuum, typically 50 in each spectrum. The point ordinates $`y_i`$ are then corrected by an amount, such that the mean of the differences between the new ordinate and all the flux values within $`1\sigma `$ (defined by the first fit) in a 6 Å wide spectral window around $`x_i`$ equals zero. The corrected $`(x_i,y_i)`$ are used to perform a new fit. This procedure is iterated until it converges to a reasonable continuum, provided enough featureless regions are available in the data. The fit routine is robust in the sense that different start values always lead to the same continuum, but the spectral windows need to be carefully chosen, especially in the B spectrum. The final continuum is shown as dotted line in Fig. 1. Note that we have not attempted to model the QSO intrinsic continuum of B, which is evidently distorted by the BALs. Division of the flux by this continuum results in the normalized spectra of HS 1216+5032 A and B. ## 3 Absorption line parameters and identifications Tables 1 and 2 list absorption lines found at the $`3\sigma `$ level in HS 1216+5032 A and B, respectively. Lines in the normalized spectra were fitted with Gaussian profiles without constraints in central wavelength, width or amplitude. In the case of obvious blends two or more Gaussian components were fitted simultaneously. The fit routine attempts to minimize $`\chi ^2`$ between model and data, considering $`1\sigma `$ flux errors (flat-fielding, background subtraction and photon statistics noise); but the uncertainties introduced by the continuum placement are not included. An inherent feature of $`\chi ^2`$ minimization is the non-uniqueness of the solution due to the eventual presence of more than one minimum in the parameter space. In those cases, the fit can lead not only to wrong parameter estimations but also to underestimated parameter errors. To handle with this problem, instead of defining a rejection parameter we simply performed several fit trials taking different wavelength ranges to see how robust a particular solution was. In general, there were no significant changes in the resulting parameters due to a different choice of the fitting interval, thus giving us confidence on the results. However, some fits yield $`\chi _\nu ^2<1`$. Assuming that the $`1\sigma `$ flux errors are not overestimated, such behaviour can occur by chance, especially when too few pixels are considered in the fit. The equivalent-width error estimates are generally larger than the propagated error if the flux values are integrated along a line. This is because the former errors consider the correlation between all three free parameters. This is not the case for multicomponent fits, where the errors in equivalent width may be overestimated since the fit algorithm does not consider that the strengths and widths of neighbour lines in blends are anti-correlated. Concerning the line widths, note that several FWHM values in Tables 1 and 2 are slightly larger than the nominal width of the line spread function (LSF), $`2`$ Å. This is caused by the difficulty in resolving absorption structures separated by less than $`200`$ km s<sup>-1</sup>, since single-line fits of closely lying line complexes (blends) will inevitably lead to large widths. On the other hand, in few cases the measured lines are narrower than the LSF. While this might partly be caused by a too low placement of the continuum near emission lines and BAL troughs in the B spectrum, it is certainly not the case for all the narrow lines in the A spectrum. In the latter case fitting weak lines also leads to low FWHM values (with large associated errors), thus reflecting the limitations of the $`\chi ^2`$ minimization using too few pixels. BAL profiles in the B spectrum have been fitted with multicomponent Gaussians to better constrain their equivalent widths, but these fits do not necessarily represent the actual nature of the BAL profiles (see § 5). For consistency, Tables 1 and 2 include the Gaussian parameters of BAL profiles. Absorption lines were identified manually, using vacuum wavelengths and oscillator strengths taken from Verner et al. (Verner (1994)). The identification of lines is not easy because of the low resolution and the lack of optical data; but it is especially difficult in the B spectrum because narrow intergalactic lines are hidden among the BALs. We looked first for interstellar absorption lines in both spectra. The Mg ii $`\lambda 2796,2803`$ doublet and the Fe ii lines at 2586 Å and 2600 Å are detected in both spectra; the Fe ii lines at 2344 Å and 2382 Å are detected at $`3\sigma `$ only in the A spectrum (while in the B spectrum Fe ii $`\lambda 2382`$ would appear at the position of BAL C iii $`\lambda 977`$). The next step was to search for C iv systems and their associated Ly$`\alpha `$ and prominent metal lines. In addition, two Mg ii systems at $`z=0.04`$ (§ 4.2) and $`1.14`$ could be identified in the spectrum of B through the positions of the doublet lines and their relative strengths. Due to their secure identification, C iv $`\lambda 1548,1550`$ and Mg ii $`\lambda 2796,2803`$ lines with $`\mathrm{SNR}<3`$ have been retained in Tables 1 and 2, which otherwise contain only lines with $`\mathrm{SNR}>3`$. Ly$`\alpha `$ lines with redshifts of $`1.16<z<z_e`$ were identified through the corresponding Ly$`\beta `$ and Ly$`\gamma `$ lines, if present. Lines A45, A46, B37, and B38 mimic a C iv doublet at $`z=0.88`$, but since no significant absorption by H i is detected at this redshift, they have been counted as Ly$`\alpha `$ lines. The remaining lines have been considered to be Ly$`\alpha `$ lines and will be discussed in § 6. ## 4 Metal absorption systems in HS 1216+5032 A and B We now give a description of the most remarkable metal absorption systems observed in HS 1216+5032 A and B: a Mg ii system at $`z=0.04`$, and a strong C iv system at $`z=0.72`$. ### 4.1 The C iv systems at $`z=\mathit{0.72}`$ in HS 1216+5032 A and B Three strong C iv systems at $`z=0.717`$, $`0.721`$ and $`0.726`$ are identified through the $`\lambda \lambda 1548,1550`$ doublets at $`\lambda 2660`$ Å in both spectra of HS 1216+5032. Although some of the C iv lines are under the $`3\sigma `$ significance level, the well-determined redshifts of the doublet components make their identification unambiguous; consequently, all six C iv lines appear in Tables 1 and 2. The rest frame equivalent widths of the $`\lambda 1548`$ lines range between $`0.36`$ and $`1.03`$ Å in the A spectrum, and between 0.72 and 1.19 Å in B. The large column densities implied by these line strengths, make this system a firm candidate for a Lyman-limit system (Sargent, Steidel & Boksenberg Sargent (1989)). Also, the large line widths in A and B, FWHM $`>2`$ Å, suggest that these C iv systems will reveal several narrower components at higher resolution. No further metal lines are found at these redshifts. Line A13 could be identified with Si iv $`\lambda 1393`$ at $`z=0.717`$, but two strong Ly$`\alpha `$ lines at $`\lambda =2410`$ Å make the detection of the second doublet line impossible. No transitions by low-ionization species (e. g., C ii $`\lambda 1334`$) are detected. In the low-resolution optical spectrum of HS 1216+5032 B (see Fig. in Hagen et al. Hagen (1996)), a significant absorption feature at $`\lambda =4816`$ Å could tentatively be identified with a Mg ii $`\lambda 2796,2803`$ doublet associated to these C iv systems, but no absorption feature is seen at this wavelength in the optical spectrum of A. The small velocity differences of $`\mathrm{\Delta }v_{\mathrm{A}\mathrm{B}}68\pm 28`$ km s<sup>-1</sup>, $`52\pm 31`$ km s<sup>-1</sup>, and $`116\pm 30`$ km s<sup>-1</sup> at $`z=0.717`$, $`0.721`$ and $`0.726`$, respectively, between the C iv lines in A and the corresponding ones in B suggest strongly that these C iv systems are physically associated. Therefore, the large velocity span of roughly $`1500`$ km s<sup>-1</sup> along each LOS can only be explained if the gas is virialized by a cluster of galaxies. The virial mass within the radius given by the half separation between the LOSs is derived to be $`10^{13}`$ M. Consequently, if C iv systems arise in the extended halos of galaxies (Bergeron & Boisse Bergeron (1991)), the present data gives evidence that the C iv absorbers at $`z=0.72`$ in HS 1216+5032 A and B arise either in the highly ionized intra-group gas of a galaxy cluster, or in small structures associated with the cluster. The whole absorption complex shows an asymmetry in the sense that the weakest system in spectrum A has the same redshift as the strongest in B: the ratio of the $`\lambda 1548`$ line’s equivalent width in A to that in B varies between 1.4 and 3.3. If C iv systems in A and B with similar redshifts are physically associated, then the different line strengths have direct implications for the size of these absorbers. This is because such gas inhomogeneities imply that the LOSs sample the clouds on spatial scales similar to the transverse cloud sizes. The interpretation remains valid if the lines in the present FOS spectra result from many unresolved velocity components, because in that case stronger lines result from more narrow components than weaker lines do, a picture that is still compatible with gas inhomogeneities. All these three C iv absorption systems should therefore arise in clouds with characteristic transverse lengths of $`75h_{50}^1`$kpc, larger than the statistical lower limits of $`30h_{50}^1`$kpc derived by Smette et al. (Smette (1995)) for C iv absorbers at $`z1.5`$. Alternatively, it is still possible that the C iv absorption is correlated in redshift, but occurs in distinct, separated structures. At higher resolution, Rauch (Rauch1 (1997)) has shown that in gas associated with metal systems density gradients on sub-kpc scales are not uncommon. The logical conclusion would be that C iv absorbers are composed of a large number of small cloudlets. However, if that is the case of the $`z=0.72`$ absorbers, the correlation in redshift between lines in A and B is difficult to explain without invoking cloudlets aligned along a filamentary or sheet-like structure. ### 4.2 A Mg ii System at $`z=\mathit{0.04}`$ in HS 1216+5032 B? There is a strong and blended feature at $`\lambda 2920`$ Å in the B spectrum for which we do not find any plausible identification with other observed metal systems. One possible identification is absorption by two Mg ii $`\lambda \lambda 2796,2803`$ doublets at $`z=0.043`$ and $`z=0.044`$. Although the Gaussian profile fit is not able to resolve the single lines in the red trough of the absorption feature, the identification is supported by a good match between the line positions and profiles of the doublet. The red wing of the whole absorption feature could be blended with a Ly$`\alpha `$ line because it coincides quite well with a strong Ly$`\alpha `$ line at $`z=1.4`$ in A (A47). However, no Fe ii lines are detected at this redshift, nor further transitions by low-ionization species. The Mg i $`\lambda 2852`$ line is possibly present, but the match in wavelength with line B50, $`\mathrm{\Delta }\lambda 1.4`$ Å, is not very good. Under the Mg ii-hypothesis, the total rest-frame equivalent width of the stronger Mg ii doublet component would be $`W_0=3.97`$ Å, which is somewhat larger than typical values found in gas associated with damped Ly$`\alpha `$ (DLA) systems at high redshift (Lu et al. Lu (1996)) or in the Milky Way (Savage et al. Savage (1993)). This could be explained if the lines are indeed made up of several narrower components, as usually seen in DLA systems. The redshift difference between both Mg ii systems implies a velocity span of roughly 300 km s<sup>-1</sup>, also typical of DLA systems. Therefore, there is some evidence that these Mg ii lines might be associated with a DLA system at $`z=0.04`$. Regardless of the DLA-system interpretation, however, even the absence of Fe ii and – possibly – also Mg i lines associated with Mg ii of such a strength is difficult to explain considering the incidence of the former ions in Mg ii-selected samples (e.g., Bergeron & Stasinska Bergeron2 (1986); Steidel & Sargent Steidel (1992)). Consequently, an alternative identification of the $`\lambda 2920`$ Å feature as H i $`\lambda 1215`$ BAL at $`v\mathrm{6\hspace{0.17em}000}`$ km s<sup>-1</sup> (see §5 and Figure 2) must be considered as well. The absence of a corresponding H i $`\lambda 1025`$ BAL profile, however, is also in this case remarkable, but could be explained if the absorber does not completely cover the continuum source (see below). ## 5 The BAL systems in HS 1216+5032 B: Some Qualitative Inferences Three BAL systems are observed in the UV spectrum of HS 1216+5032 B (see Fig. 2, where the system redshifts have been arbitrarily numbered 1, 2, and 3). Absorption by H i, C iii, N iii, and possibly S iv is observed in at least two of the systems, while O vi and N v are present in all three systems. C ii is observed only in system 2, and N iii in systems 2 and 3. Many features distinguish these systems from the most commonly observed BAL systems (e.g., Turnshek et al. Turnshek (1996)): (1) the lines are particularly weak and several line profiles are not distorted by/or blended with other BALs; (2) the maximum outflow velocity $`v\mathrm{5\hspace{0.17em}000}`$ km s<sup>-1</sup> is small compared with typical BALQSOs, which exhibit terminal velocities of several $`10^4`$ km s<sup>-1</sup>(Turnshek Turnshek1 (1984)); (3) C ii is present (see Wampler et al. Wampler (1995) for other BAL QSO spectrum with singly-ionized species; Arav et al. 1999b ); (4) the strength of H i in systems 2 and 3 decreases more slowly at the red edge of the troughs (Turnshek Turnshek1 (1984)). Also remarkable is the undulating shape of the absorbed continuum for $`\lambda <2700`$ Å (see Fig. 1). The position of the flux depressions coincides quite well with the blue wing of expected emission lines by C iii, N iii, Ly$`\beta `$, and O vi. This coincidence seems to suggest that the absorption is indeed dominated by very broad line components, which determine the continuum shape, superimposed to the narrower components, here labeled as systems 1, 2 and 3. The same effect, although less remarkable, is observed for S iv and S vi, thus giving evidence for absorption by these ions associated with the BAL phenomenon (also reported for another QSO by Arav et al. 1999a ). It is customary to define BALs as a continuous absorption with outflow velocities larger than $`\mathrm{3\hspace{0.17em}000}`$ km s<sup>-1</sup> from the emission redshift (Weymann et al. Weymann2 (1991)) to make a distinction between BAL systems and “associated systems”. However, as pointed out by Arav et al. (1999b ), such definition does not hold any physical meaning. In our case, systems 1 and 2 should be then classified as associated systems but, due to our poor resolution, it is not possible to establish whether the line profiles are produced by continuous absorption or whether they are made of several narrower velocity components (as observed in associated systems). The metal lines in these systems (e.g. the N v and O vi doublet lines) are relatively narrow and their widths could be dominated by the instrumental profile (for H i in system 2, however, the situation is less clear). Therefore, the classification of systems 1 and 2 as BAL systems must be considered solely instrumental. High-resolution spectra of BAL QSOs (Hamann et al. Hamman (1997)) show that BAL profiles do not necessarily result from an ensemble of discrete narrow, unresolved lines, but as part of a mixture of a continuously accelerated outflow and overlapping narrow components with different non-thermal velocity dispersions. For the systems in HS 1216+5032, it can be assumed that a similar mixture of line widths is present, with the broad components dominating over the narrow ones. In that case, these line profiles should not look so different at higher resolution (with exception, maybe, of systems 1 and 2). Ionization models have shown that BAL clouds span a range of densities and/or distances from the ionizing source (Turnshek et al. Turnshek (1996); Hamann et al. Hamman1 (1995)). Fig. 2 shows that, while the high-ionization species N v and O vi are present in all three systems in HS 1216+5032, H i, C ii and doubly-ionized species are absent in system 1. A possible explanation is that the ionization conditions could change with outflow velocity, with higher ionization level for redshifts closer to $`z_e`$. However, a quantitative study of the ionization conditions in these BAL clouds with the present data is made difficult by our inability to reliably establish the continuum level at the BAL troughs and thus determine column densities (besides the fact that nonblack saturation of the line profiles might be present; see Arav et al. 1999a ). Nevertheless, a more quantitative study of the BAL phenomenon in HS 1216+5032 B should be possible using higher resolution HST spectra to better estimate the QSO continuum and to identify spurious lines at the BAL troughs. In particular, BAL system 2 can be studied in more detail because it shows more ions and less contamination by narrow absorption lines at the lines lying in the Ly$`\alpha `$ forest. On the other hand, alone medium-resolution optical spectra should considerably improve our knowledge of the BAL systems in HS 1216+5032 through the line profiles of C iv, C iii, and possibly Mg ii, which we suspect to be present given the presence of C ii. The width of the trough in system 2 is sufficiently small to study nonblack saturation effects via doublet ratios provided the emission line profiles can be determined (e.g., at the position of the C iv BAL). Thus, corrected column densities for system 2 appear as very suitable to reliably examine photoionization models because the radiation fields can be constrained by the presence of low-ionization species. ## 6 Ly$`\alpha `$ absorption systems We now describe the Ly$`\alpha `$ forest lines observed in the spectra of HS 1216+5032 A and B. The projected separation between LOSs samples well the expected cloud sizes of hundreds of kpc if the QSO pair is real (see next subsection). However, the effective redshift path where Ly$`\alpha `$ lines can in principle be detected is blocked out by the broad absorption lines in the B spectrum, so the line samples will be small. ### 6.1 Lensed or Physical Pair? The nature of the double images in HS 1216+5032 is not clearly established yet. Maybe the strongest arguments favoring HS 1216+5032 AB as a physical-pair instead of a gravitational lens origin are: (1) BALs are observed only in the spectrum of the B image;<sup>2</sup><sup>2</sup>2 Curiously, another QSO pair separated by $`9\mathrm{}`$, Q1343+266, also shows BALs in only one spectrum. On the basis of the irregular flux ratio in the optical range, Crotts et al. (Crotts (1994)) find evidence that this is a genuine QSO pair. From the optical spectra of HS 1216+5032 AB (Hagen et al. Hagen (1996)), however, a similar argument is less clear. (2) no extended, luminous object is detected between (or even close to) the QSO images down to $`m_R=22.5`$; and (3) gravitationally lensed QSO pairs normally have image separations $`\theta 3\mathrm{}`$ (e.g., Kochanek, Falco & Muñoz Kochanek (1999)). In addition, the absence of a trend for equivalent widths of intervening Ly$`\alpha `$ lines common to both spectra to become similar at redshifts closer to $`z_e`$ (see Fig. 4 and Table 3) also argues for a physical-pair; however, this argument is weak as the sample of common lines is too small. Unfortunately, for reasons given in § 2.2, it is not possible to establish whether there are differences between A and B in the emission line redshifts and line shapes in the spectral region covered by the FOS. The intriguing point here is that two $`zz_e`$ Ly$`\alpha `$ systems are detected in both spectra at almost identical redshifts to within $`90`$ km s<sup>-1</sup> (see § 6.6 for more details). Under the physical-pair hypothesis, these absorbers should extend over transverse scales greater than $`80h_{50}^1`$kpc. If, on the other hand, HS 1216+5032 AB is a gravitationally lensed pair, the detection of these systems in both spectra can easily be explained by the fact that the separation between light paths approaches zero. In consequence, the gravitational lens nature of HS 1216+5032 cannot be yet ruled out. Let us note that in a simple lens geometry the virial mass implied by the velocity span of the C iv absorbers at $`z=0.72`$ (see § 4.1) is consistent with the deflector mass required to produce an angular separation between QSO images of $`\theta =9\mathrm{}`$. In such a model the separation between light paths at the source position is derived to be $`\theta _s=17`$″. Therefore, confirmation of HS 1216+5032 as a gravitationally lensed QSO would have dramatic consequences for our understanding of the BAL phenomenon. As pointed out by Hagen et al. (Hagen (1996)), if HS 1216+5032 AB were indeed the mirror images of one unique QSO, the coverage fraction of BAL clouds would have to be very small. Incomplete coverage of the continuum source or the BLR in BAL clouds have been confirmed using high-resolution spectra (Hamann et al. Hamman (1997); Barlow & Sargent Barlow (1997)). In fact, the cloud sizes derived from BAL variability could be $`10^4`$ times smaller than the distances to the source derived from photoionization models (Hamann et al. Hamman1 (1995)), implying subtended angles not much larger than $`\theta _s`$. Therefore, the scenario in which only one light path crosses the BAL clouds is not unrealistic, and the non-BAL spectrum of HS 1216+5032 A can also be considered if the QSO pair is lensed.<sup>3</sup><sup>3</sup>3Let us mention, however, that in the lensed-pair hypothesis the fact that BAL absorption is produced by many clouds (Hamann et al. Hamman (1997)) lessens considerably the probability that LOS A does not encounter neither such clouds. In summary, there are reasons to believe that HS 1216+5032 is a binary QSO, but a gravitational lens origin of the double images cannot yet be excluded. A definitive assessment of this issue must await medium resolution optical spectra to better derive emission redshifts and to establish differences in the continuum or emission lines. Alternatively, deep infrared or radio observations would help determining the nature of the double QSO images. The non-detection of the lensing galaxy/cluster would be a strong proof for the binary QSO hypothesis, since wide separation ($`\theta >3\mathrm{}`$) “dark lenses” enter in conflict with current models of structure formation (Kochanek, Falco & Muñoz Kochanek (1999)), which predict few wide separation gravitational lenses (e.g., Kochanek Kochanek1 (1995)). Throughout this section we will assume that HS 1216+5032 AB is a physical pair. The proper separation $`S(z)`$ between LOSs obeys then the relation $`S=\theta D_{\mathrm{oc}}`$, where $`\theta `$ is the angular separation between the A and B QSO images, and $`D_{\mathrm{oc}}`$ is the angular-diameter distance between the observer and a cloud at redshift $`z`$. For the redshift range of interest $`S80`$ $`h_{50}^1`$ kpc. ### 6.2 Line List Table 3 lists Ly$`\alpha `$ lines detected at the $`3\sigma `$ level in HS 1216+5032 A and B. Figure 3 shows $`3\sigma `$ detection thresholds as a function of redshift. Out of 36 lines in the spectrum of A, a total of 23 lines within $`z=0.85`$ to $`1.43`$ have $`W_0(\mathrm{A})0.24`$ Å. This redshift interval excludes lines within $`\mathrm{3\hspace{0.17em}000}`$ km s<sup>-1</sup> from the QSO redshift to avoid the “proximity effect” (Bechtold Bechtold (1994)). The derived number density is $`\mathrm{log}(dN/dz)=1.58_{0.10}^{+0.09}`$ at $`\mathrm{log}(1+z)=0.33\pm 0.06`$, in very good agreement with the results of the HST Key Project (Weymann et al. Weymann1 (1998)). The scarce sample of lines in B is clearly explained by the shorter redshift path allowed by the BAL profiles. ### 6.3 Definition of Line Samples To define the number of coincidences and anti-coincidences we have selected from all 36 Ly$`\alpha `$ lines observed in A (the spectrum with better signal-to-noise) those ones (1) at $`z1.43`$; (2) not associated with metal lines; (3) at wavelengths not covered by the BAL troughs in B; and (4) at wavelengths where $`3\sigma `$ detection limits in B are lower than the measured equivalent width of the line in A. Lines in B that had $`\mathrm{SNR}>3`$ and fulfilled criteria (1) to (3) were also selected. In some cases, lines in A have a corresponding absorption feature at the same wavelength in B, but at a too low significance level to unambiguously designate the line pair as a coincidence. These lines in A were excluded (A26, A36, A44). Notice that criterion (3) automatically prevents comparing line equivalent widths for which the uncertainties introduced by the placement of the B continuum are large. The final sample of Ly$`\alpha `$ lines suitable for this study is composed of $`N=11`$ redshift systems within $`1.16<z<1.43`$, a range not much smaller than the one allowed by conditions (1) to (4). Out of this number, $`N_\mathrm{C}=6`$ systems show Ly$`\alpha `$ lines in both spectra, and $`N_\mathrm{A}=5`$ lines are detected in only one spectrum. The rest-frame equivalent widths range from 0.14 to 0.76 Å in A, and from 0.23 to 1.16 Å in B. This sample is hereafter called “full sample”. Coincident lines are shown in Figure 4, where the thick line represents the flux of A and the zero velocity point corresponds to the redshift of the A line. Only the first six panels show lines in the “full sample”; the remaining lines arise in Ly$`\alpha `$ clouds likely to be influenced by the QSO flux. All coincident lines are separated by less than 100 km s<sup>-1</sup>. Anti-coincident lines are all uniquely defined, as there are no corresponding lines in the other spectrum within several hundred km s<sup>-1</sup>. However, there are two exceptions that must be pointed out. The first one is line 29 in B, separated by $`198`$ km s<sup>-1</sup> from the nearest Ly$`\alpha `$ line in A (37). These lines in A and B have been counted as anti-coincidences, though we are suspicious of this interpretation as the line profiles suggest line B29 is blended with an absorption feature seen in both spectra. The second is another of the anti-coincidences, line B50, which could alternatively be identified with Mg i at $`z=0.04`$. These cases will introduce unavoidable uncertainties in the results presented here. Let us recall that the exclusion of any one line from the samples would lead to significantly different results for $`R_\mathrm{c}`$. This illustrates the real uncertainties dominating simulations with such a small number of observed lines. Moreover, the samples are limited by the large redshift path blocked out by the BAL troughs in B, so there must be a considerable loss of information. A line significance level of $`5`$ for lines in A has been chosen to perform the likelihood analysis described in the next subsection . This selection automatically excludes two of the redshift systems in the full sample and defines the following sub-samples: 1. Strong-line sample (sample 1): lines for which SNR(A) $`>5`$, SNR(B) $`>3`$, and $`W_0>0.32`$ Å. This selection yields $`N_\mathrm{C}=2`$ coincidences and $`N_\mathrm{A}=4`$ anti-coincidences. 2. Strong+weak line sample (sample 2): same as previous sample but with $`W_0>0.17`$ Å. This selection yields $`N_\mathrm{C}=4`$ and $`N_\mathrm{A}=5`$. ### 6.4 Maximum Likelihood Analysis In the following we discuss a likelihood analysis on cloud sizes in front of HS 1216+5032. The technique is based on the definition of a likelihood function $``$ that gives the probability of getting the observed number of coincidences and anti-coincidences. Evidently, such a function must depend on the shape of the absorber. Here we concentrate on two possible cloud geometries for which $``$ can be derived analytically: spheres and cylinders. For coherent spherical clouds, the probability that one cloud at redshift $`z`$ is intersected by one LOS is given by McGill (McGill (1990)) as: $$\varphi (X)=\frac{2}{\pi }\left\{\mathrm{arccos}\left[X(z)\right]X(z)\sqrt{1X(z)^2}\right\},$$ (1) where $`X(z)S(z)/2R_\mathrm{c}`$ and $`R_\mathrm{c}`$ is the cloud radius. The samples consider only redshifts where at least one LOS intersects a cloud (otherwise one should make assumptions on the cloud distribution along the LOS). Therefore, one must compute the probability that, given that a line appears in one spectrum, a line will appear in the other spectrum. This probability is given by Dinshaw et al. (Dinshaw2 (1997)) as: $$\psi (X)=\frac{\varphi }{2\varphi }.$$ (2) Finally, the probability of getting the observed number of coincidences and anti-coincidences is given by the product $$(R_\mathrm{c})=\underset{i}{}\psi \left[X(z_i)\right]\underset{j}{}\left\{1\psi \left[X(z_j)\right]\right\},$$ (3) where the indexes $`i`$ and $`j`$ number the coincidences and anti-coincidences, respectively. Note that if there is at least one anti-coincidence, then $``$ has a maximum value. Otherwise it grows monotonically. The solid curve in Fig. 6 shows the results of the likelihood function $`(R_c)`$ normalized to its peak intensity for the various samples. The figure also shows the cumulative distribution of $`(R_c)`$. The estimated $`2\sigma `$ limits on cloud diameters for HS 1216+5032 – derived from the cumulative distribution – are listed in Table 4. The peak intensity of $``$ is used to find the most probably radii, i.e., $`96`$ $`h_{50}^1`$ kpc for the strong line sample and $`128`$ $`h_{50}^1`$ kpc for the strong+weak line sample. From the cumulative distribution, the respective median values are $`145`$ and $`173`$ $`h_{50}^1`$ kpc. In both cases, the derived sizes are larger if weaker lines are used. This result provides evidence that Ly$`\alpha `$ clouds must have a smooth density distribution. In a second model, let us suppose that Ly$`\alpha `$ lines occur in filamentary structures lying perpendicular to the LOSs. Such structures can be idealized as cylinders with a radius-to-length ratio $`1`$. For cylindric clouds, the probability that one LOS intersects a cloud at redshift $`z`$ given that the other LOS already does, reads (A. Smette, private communication): $$\varphi (X)=\{\begin{array}{cc}\frac{2}{\pi }\{\mathrm{arcsin}\left[X(z)^1\right]\hfill & \\ +\sqrt{X(z)^21}X(z)\}\hfill & \mathrm{for}X(z)>1;\hfill \\ & \\ 1\frac{2}{\pi }X(z)\hfill & \mathrm{otherwise},\hfill \end{array}$$ (4) where $`X(z)S(z)/2R_\mathrm{c}`$ and $`R_\mathrm{c}`$ is the radius of the cylinder. The dotted curve in Fig. 6 shows the results of the likelihood function $`(R_c)`$ normalized to its peak intensity and the cumulative distribution for the various samples. Most probably cylinder radii are $`49`$, $`65`$, and $`85`$ $`h_{50}^1`$ kpc for the strong, strong+weak and full line samples, respectively. The respective median values are $`70`$, $`87`$ and $`115`$ $`h_{50}^1`$ kpc. Two sigma bounds are displayed in Table 4. Again, the derived sizes are larger if weaker lines are used, providing evidence that the density distribution in Ly$`\alpha `$ clouds must be smooth (see, e.g., Monier et al. Monier (1999)). These size estimates are almost $`50`$% lower than for spherical absorbers. This can be explained by the fact that, in general, elongated structures can more easily reproduce the observed number of coincidences for a given radius than a sphere; in fact, the probability $`\varphi (R_c)`$ for spherical absorbers vanishes for $`R_c<S/2`$, while for cylinders $`\varphi >0`$ for all nonzero values of $`R_c`$. The lengths of such structures are much larger than these values, but undefined in the model. ### 6.5 Equivalent Widths The sample of Ly$`\alpha `$ lines common to both spectra gives model-independent information about the size scales of the absorbers. Figure 7 shows the rest-frame equivalent widths of lines common to both spectra and their $`1\sigma `$ errors, where the dashed straight line has a slope of unity. Only lines without associated metal lines are included. Note that this selection yields line-pairs with $`\lambda >2620`$ Å, so the influence of uncertainties in the continuum placement (e.g. due to the BAL troughs or to spectral regions with high noise level) over the equivalent-width estimates is minimized. We have included one line pair within 3 000 km s<sup>-1</sup> from the QSO’s redshift (A52,B49) but not the “associated system”. The dotted crosses represent line pairs for which the B line has $`2.58<\mathrm{SNR}<3`$, i.e., those ones labeled with footnote 5 in Table 3. Additionally, to illustrate the significance of the non-detections we have also included lines detected in only one spectrum, plotted against the corresponding $`3\sigma `$ detection limit in the other spectrum. Most of the coincident lines are stronger in the B spectrum, which is explained by the difficulty in detecting lines in B (an additional consequence of this is that most anti-coincidences are lines in B). It can be seen that there are some significant deviations from $`\mathrm{\Delta }W_0=0`$, even excluding the line pair within 3 000 km s<sup>-1</sup>. For the latter such deviation would be expected, since the QSOs themselves might be an important ionizing agent in clouds next to the QSOs.<sup>4</sup><sup>4</sup>4Indeed, if the QSO radiation field contributed significantly to ionize these clouds, one would expect $`W(\mathrm{A})<W(\mathrm{B})`$ for these lines, because QSO A is more luminous than B. Such difference in equivalent width is, however, not observed. The equivalent width differences $`\mathrm{\Delta }W_0=|W_0(\mathrm{A})W_0(\mathrm{B})|`$ show no correlation with $`\mathrm{\Delta }v`$. Instead, $`\mathrm{\Delta }W_0`$ and max$`(W_0(\mathrm{A}),W_0(\mathrm{B}))`$ seem to be correlated (see Fig. 8), with larger equivalent width differences for larger equivalent widths (e.g. Fang et al. Fang (1996)). Both effects, although not statistically significant, suggest coherent structures, i.e., no “cloudlets” at similar redshifts (Charlton et al. Charlton (1995)). In conclusion, LOS A and B must sample coherent clouds showing small – but otherwise significant – density gradients on spatial scales of $`80h_{50}^1`$kpc, the linear separation between LOSs in this redshift range. ### 6.6 The $`z_az_e`$ Ly$`\alpha `$ systems in HS 1216+5032 A and B Two $`z_az_e`$ Ly$`\alpha `$ systems are observed in the spectra of both HS 1216+5032 A and B, through the partially resolved Ly$`\alpha `$ lines A53, A54, A55, and B51 and B52 (see Fig. 4). The lines in A and B differ by $`\mathrm{\Delta }v_{\mathrm{A53}\mathrm{B51}}=51`$ km s<sup>-1</sup>and $`\mathrm{\Delta }v_{\mathrm{A54}\mathrm{B52}}=87`$ km s<sup>-1</sup>. The system in A is blueshifted by 122 km s<sup>-1</sup> relative to $`z_e(\mathrm{A})`$, while the system in B is redshifted by 280 km s<sup>-1</sup> relative to $`z_e(\mathrm{B})`$ (but see § 2.2). Line A55 has no clear counterpart in B. No metal lines are found associated with these systems. The total equivalent width of the A lines is larger than in B by a factor of 2.1. The absence of highly ionized species such as N v and O vi suggests that these systems are probably not physically associated with the QSOs (e.g., Turnshek Turnshek1 (1984); Petitjean Petitjean1 (1994)). Furthermore, the small velocity differences between lines in A and B, and the fact that both systems have two line components well correlated in redshift, strongly suggest they arise in the same absorber (see also Petitjean et al. Petitjean (1998); Shaver & Robertson Shaver (1983)). If this is true, the different Ly$`\alpha `$ line strengths between A and B imply either characteristic transverse sizes of $`80h_{50}^1`$kpc or maybe even larger clouds with ionization conditions that change on such scales. The size scales are compatible with the idea that these systems arise, for example, in the very extended halo of an intervening galaxy (Weymann et al. Weymann (1979)), or in the intra-group gas of the QSO host galaxies. If the radiation field of the QSOs is the main ionization source in the clouds giving rise to these systems (assuming the QSO pair is real), then the systems in B should be less ionized than the A ones. This is because QSO A is intrinsically more luminous than B in the ionizing continuum. Consequently, the different line strengths in A and B can also be explained if the LOSs cross regions of similar density, with LOS B probing regions with more neutral gas. Of course, the latter case does not rule out the intervening nature of these systems. ### 6.7 Discussion Fang et al. (Fang (1996)) have pointed out that there seems to be a trend of larger estimated cloud sizes with increasing LOS separation $`S`$ (see their Fig. 5). If such a trend were real, it would imply that the scenario of uniform-sized spherical clouds is too idealized. Their data, however, are insufficient to discern whether the effect is due to non-uniform cloud size or simply non-spherical cloud geometry. The results presented in our work (slightly modified by defining a sample of lines with $`W_0>0.40`$ Å to be consistent with these authors) fit this tendency well because they add a new measurement of relatively small cloud sizes at small LOS separation. We note, however, that D’Odorico et al. (D'Odorico (1998)), who use a larger database and an improved statistical approach, find no correlation between cloud size and LOS separation. Clearly, more QSO pairs are needed that span a range in LOS separations to confirm or discard the size/separation correlation. Here we propose that the notion of a single population of uniform-sized clouds must be revised. In fact, if Ly$`\alpha `$ clouds span a range of sizes between, say, a few hundred kpc to a few Mpc, then LOSs to QSO pairs with arcminute separations would be crossing not only huge and coherent structures, but also smaller clouds correlated in redshift. Further evidence for a more complicated scenario than usually assumed comes from cosmological hydrodynamic simulations made for $`z=3`$, which show entities of non-uniform sizes grouped along filamentary structures (e.g., Cen & Simcoe Cen (1997)). However, we recall that the situation at lower redshift can be different if such structures do evolve in size. Size evolution of Ly$`\alpha `$ absorbers has not yet been observed, in part because the number of adjacent QSOs with suitable angular separations is not statistically significant, but also due to the scarce number of observations at low redshift. In particular, the present data on HS 1216+5032 ($`<z>=1.3`$) do not confirm the suggestion (Fang et al. Fang (1996); Dinshaw et al. Dinshaw1 (1998)) that cloud sizes increase with decreasing redshift. This result is in variance with the findings by D’Odorico et al. (D'Odorico (1998)). Assuming that the Ly$`\alpha `$ absorbers are filamentary structures – modelled as cylinders of infinite length – lying perpendicular to the LOSs, our likelihood analysis leads to almost $`50\%`$ smaller transverse dimensions than for spherical clouds. Unfortunately, the method presented here is not capable of distinguishing between cylinders and spheres. Moreover, the information provided by two LOSs might be insufficient to make such a distinction possible, so one should use triply imaged QSOs (Crotts & Fang Crotts1 (1998)) or, even better, QSO groups (Monier et al. Monier (1999)). It is worth saying, however, that flattened structures are more able to simultaneously reproduce the requirements of neutral gas density from photoionization models and the transverse scale lengths derived from double LOSs than spherical absorbers do. Such photoionization models lead to structures with thickness-to-length ratios of $`1/30`$ (Rauch & Haehnelt Rauch4 (1995)). Indeed, hydrodynamical simulations in the context of hierarchical structure formation have shown that at $`z3`$, high density gas regions (producing $`W_0>0.3`$ Å Ly$`\alpha `$ absorption lines) are connected by filamentary and sheet-like structures roughly $`10^6`$ times less dense than the embedded condensations. The filaments seem to evolve slowly and still fill the Universe at $`z1`$ (Davé et al. Dave (1999); Riediger, Petitjean & Mücket Riediger (1998)), giving rise to the majority of strong Ly$`\alpha `$ lines. If this picture is correct, the requirement of different cloud populations might not be necessary to explain current observations. New high-resolution ultraviolet observations of QSO pairs are needed. Despite technical difficulties (close QSO pairs tend to have such different observed fluxes that high dispersion spectroscopy normally leads to at least one spectrum with high noise levels) they should improve considerably our knowledge of the Ly$`\alpha `$-cloud geometry by (1) analyzing absorbers that produce only weak ($`W_00.01`$ Å) lines, i.e., those surely not associated with low-brightness galaxies, and (2) testing models that consider column density distributions. ## 7 Summary We have presented HST FOS spectra of the QSO pair HS 1216+5032 AB. Our results are summarized as follows: 1. Metal systems. Three strong and complex C iv absorption systems are observed in both HST spectra at $`z0.72`$. The velocity span along the LOSs, $`\mathrm{\Delta }v1500`$ km s<sup>-1</sup>, is large and suggests the gas where this C iv occurs might be associated with a cluster of galaxies. Differences in the line strengths suggest that, if these systems arise in coherent structures, they must have characteristic transverse sizes of $`75h_{50}^1`$kpc. Alternatively, the systems may be composed of a large number of cloudlets correlated in redshift. In the spectrum of B, a strong absorption feature at $`\lambda 2910`$ Å is identified with two Mg ii doublets at $`z=0.04`$. If this identification is correct, these lines could arise in a low-redshift damped Ly$`\alpha `$ system. 2. The BAL systems in B: The spectrum of QSO image B shows BAL troughs by H i, C ii, C iii, N iii, N v, O vi, and possibly S iv and S vi arising in absorption systems at outflow velocities from the QSO of up to $`5000`$ km s<sup>-1</sup>. The BAL troughs arise probably as a consequence of absorption by a mixture of broad and narrow components. 3. Ly$`\alpha `$ absorbers. Due to the redshift path blocked out by the BAL troughs in B the number of detected lines is small. Selecting lines not associated with metal lines, with $`\mathrm{\Delta }v(\mathrm{A}\mathrm{B})<100`$ km s<sup>-1</sup> and $`W_0>0.17`$ Å yields four lines common to both spectra and five lines without counterpart in the other spectrum. For $`W_0>0.32`$ Å lines the numbers are two and four, respectively. Using a maximum likelihood technique, most probably diameters for spherical clouds of $`192`$ and $`256`$ $`h_{50}^1`$ kpc are found for $`W>0.32`$ Å and $`W>0.17`$ Å lines, respectively. The $`2\sigma `$ limits derived using the cumulative distribution of the probability function are $`136<D<880`$ $`h_{50}^1`$ kpc and $`172<D<896`$ $`h_{50}^1`$ kpc for the respective samples at $`<z>=1.3`$. Assuming that the absorbers are filamentary structures lying perpendicular to the LOSs, transverse dimensions almost $`50\%`$ smaller than for spherical clouds are found. In both cases, the results of our analysis do not confirm the claim that the characteristic size of the Ly$`\alpha `$ absorbers increases with decreasing redshift. Independently of the cloud models used, we note that there are significant equivalent width differences between lines in A and B. Also, there appears to be a trend of larger equivalent width differences with increasing line strength, while no velocity differences between common lines is found. This provides evidence that the absorbers are coherent entities. The results for each line sample suggest that the absorbers must have a smooth gas density distribution, with lower density gas being more extended. ###### Acknowledgements. We are grateful to Michelle Mizuno, Lutz Wisotzki, and specially Alain Smette for their valuable comments on early drafts. We also thank Olaf Wucknitz for calculations concerning the gravitational-lens hypothesis. S. L. acknowledges support by the BMBF (DARA) under grant No. 50 OR 9905.
warning/0003/hep-th0003260.html
ar5iv
text
# Symplectic Symmetry of the Neutrino Mass and the See-Saw Mechanism ## I Introduction Neutrinos play an essential role in electroweak physics and there is a fascinating interplay between neutrino properties and various astrophysical phenomena. Recent observations of especially the zenith-angle dependence of atmospheric neutrinos , and the solar neutrino flux provide strong hints of non-zero neutrino masses and oscillations. Atmospheric and solar neutrino experiments along with direct experimental limits on the neutrino mass indicate that neutrino masses are much smaller than the charged-lepton masses. In the Standard Model neutrinos are left-handed. Hence the only Lorentz scalar pairing is the Majorana mass which is a weak isotriplet violating global lepton number conservation. Since the Standard Model has no isotriplet Higgs to couple to such a neutrino pair neutrinos are taken to be massless. Thus the neutrino mass is perhaps the most exquisite hint for physics beyond the Standard Model. In the current wisdom the Standard Model is not a fundamental, but an effective theory. Weak gauge bosons, W and Z, couple at low energies only with left-handed interactions. Neutrinos, being neutral leptons, interact only weakly at low energies. Hence the right-handed components of the neutrinos do not interact, that does not mean that they do not exist. A fundamental understanding of the origin and the magnitude of neutrino mass and mixings is lacking, however the see-saw mechanism gives a natural scheme for the smallness of the neutrino masses and provides a convenient ansatz for model building. A very nice discussion of massive neutrino scenarios beyond the standard model is given in Reference . Approaches to the neutrino mass based on symmetries are not fully exploited. Algebraic formulation of a problem could provide a check of the consistency of the formalism and may help to uncover symmetries. In particular it could be beneficial to rewrite the neutrino mass term in the Hamiltonian in terms of the elements of a Lie algebra which also generates the see-saw mechanism. The purpose of the present paper is to explore this possibility to place the neutrino mass matrix and the see-saw mechanism in the same algebraic framework. One can view the mass term in the Hamiltonian as a generalized pairing problem. For a single neutrino family (four Dirac components) the most general pairing algebra is $`SO(8)`$ (see for example ). We show that for the neutrino mass Lorentz invariance constraints this algebra to be $`Sp(4)`$. We also show that the Pauli-Gürsey transformation is an $`SU(2)_{\mathrm{PG}}`$ rotation which is embedded in the associated $`Sp(4)`$ Lie group. The see-saw Hamiltonian is generated by a particular Pauli-Gürsey transformation. The mass Hamiltonian sits in the $`Sp(4)/SU(2)_{\mathrm{PG}}\times U(1)`$ coset where the $`U(1)`$ is a chirality transformation. In the next section we introduce the $`Sp(4)`$ algebra for a single neutrino flavor and write the mass Hamiltonian in terms of its generators. In Section 3 we show that the Pauli-Gürsey transformation belongs to the associated group and obtain the see-saw mechanism as a particular Pauli-Gürsey transformation. We also discuss how the algebraic structure generalizes for the case of many flavors of neutrinos. Finally a brief discussion of our results conclude the paper. ## II The Algebraic Structure of the Neutrino Mass Matrix We shall use the chiral representation of Dirac matrices, in which $$\stackrel{}{\gamma }=\left(\begin{array}{cc}0& \stackrel{}{\sigma }\\ & \\ \stackrel{}{\sigma }& 0\end{array}\right),\gamma _5=\left(\begin{array}{cc}1& 0\\ & \\ 0& 1\end{array}\right),\gamma _0=\left(\begin{array}{cc}0& 1\\ & \\ 1& 0\end{array}\right)$$ (1) where the matrices $`\stackrel{}{\sigma }`$ are the usual $`2\times 2`$ Pauli matrices. Since we want to introduce charges we will work with the Hamiltonians instead of Hamiltonian densities. Introducing the left-and right-handed fields $$\psi _L=\frac{1}{2}(1\gamma _5)\psi ,\psi _R=\frac{1}{2}(1+\gamma _5)\psi $$ (2) the Dirac-mass term in the Hamiltonian can be written as $$H_m^D=m_Dd^3x(\overline{\psi }_L\psi _R+h.c)$$ (3) and the most general Majorana mass term as $$H_m^M=H_m^L+H_m^R=\frac{1}{2}m_Ld^3x(\overline{\psi }_L\psi _L^c+h.c)+\frac{1}{2}m_Rd^3x(\overline{\psi }_R\psi _R^c+h.c).$$ (4) In these equations $`m_D`$, $`m_L`$, and $`m_R`$ are the Dirac, left-handed and right-handed Majorana masses and the charge conjugation of $`\psi `$ field is defined by $$\psi ^C=C\overline{\psi }^T$$ (5) where $`C`$ is the charge conjugation matrix. In the chiral representation it is given by $$C=\left(\begin{array}{cc}i\sigma _2& 0\\ & \\ o& i\sigma _2\end{array}\right).$$ (6) Introducing the charges $`D_+`$ $`=`$ $`{\displaystyle d^3x(\overline{\psi }_L\psi _R)}`$ (8) $`D_{}`$ $`=`$ $`{\displaystyle d^3x(\overline{\psi }_R\psi _L)}=D_+^{}`$ (9) the canonical equal-time anti-commutation relations between the fermion fields lead to $$[D_+,D_{}]=2D_0$$ (10) where $`D_0`$ is defined as $$D_0=\frac{1}{2}d^3x(\psi _L^{}\psi _L\psi _R^{}\psi _R).$$ (11) We see that $`D_+`$, $`D_{}`$ and $`D_0`$ satisfy $`SU(2)`$ commutation relations: $$[D_+,D_{}]=2D_0,[D_0,D_+]=D_+,[D_0,D_{}]=D_{}.$$ (12) Similarly introducing the left-handed $`L_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^3x(\overline{\psi }_L\psi _L^c)}`$ (14) $`L_{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^3x(\overline{\psi _L^c}\psi _L)}=L_+^{}`$ (15) $`L_0`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle d^3x(\psi _L^{}\psi _L\psi _L\psi _L^{})}`$ (16) and the right-handed charges $`R_+`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^3x(\overline{\psi _R^c}\psi _R)}`$ (18) $`R_{}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^3x(\overline{\psi }_R\psi _R^c)}=R_+^{}`$ (19) $`R_0`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle d^3x(\psi _R\psi _R^{}\psi _R^{}\psi _R)}`$ (20) one can see that they also satisfy $`SU(2)`$ algebra structure: $$[L_+,L_{}]=2L_0,[L_0,L_+]=L_+,[L_0,L_{}]=L_{}$$ (21) $$[R_+,R_{}]=2R_0,[R_0,R_+]=R_+,[R_0,R_{}]=R_{}$$ (22) Also note that $`D_0L_0+R_0`$. It is easy to show that the $`SU(2)`$ algebras generated by the left- and right-handed charges commute: $$[L_{,+,0},R_{,+,0}]=0.$$ (23) Calculating all the commutation relations between $`D`$’s, $`L`$’s and $`R`$’s to find a closed algebra structure provides a new set of generators: $`[D_+,L_0]`$ $`=`$ $`{\displaystyle \frac{1}{2}}D_+,`$ (25) $`[D_+,L_+]`$ $`=`$ $`0,`$ (26) $`[D_+,L_{}]`$ $`=`$ $`A_+,`$ (27) $`[D_+,R_0]`$ $`=`$ $`{\displaystyle \frac{1}{2}}D_+,`$ (28) $`[D_+,R_+]`$ $`=`$ $`0,`$ (29) $`[D_+,R_{}]`$ $`=`$ $`A_{},`$ (30) $`[D_{},L_0]`$ $`=`$ $`{\displaystyle \frac{1}{2}}D_{},`$ (31) $`[D_{},L_+]`$ $`=`$ $`A_{},`$ (32) $`[D_{},L_{}]`$ $`=`$ $`0,`$ (33) $`[D_{},R_0]`$ $`=`$ $`{\displaystyle \frac{1}{2}}D_{},`$ (34) $`[D_{},R_+]`$ $`=`$ $`A_+,`$ (35) $`[D_{},R_{}]`$ $`=`$ $`0,`$ (36) where the additional generators are $`A_+={\displaystyle d^3x\left[\psi _L^TC\gamma _0\psi _R\right]},`$ (38) $`A_{}={\displaystyle d^3x\left[\psi _R^{}\gamma _0C(\psi _L^{})^T\right]}.`$ (39) The commutation relations between them give another SU(2) algebra $$[A_+,A_{}]=2(R_0L_0)2A_0,[A_+,A_0]=A_+,[A_{},A_0]=A_{}.$$ (40) Note that $`A_0`$ is proportional to the neutrino number operator. The commutation relations between $`A_+`$ and all the other generators $$[A_+,D_+]=2R_+,,[A_+,D_{}]=2L_{},$$ (41) $$[A_+,L_0]=\frac{1}{2}A_+,,[A_+,L_+]=D_+,[A_+,L_{}]=0,$$ (42) $$[A_+,R_0]=\frac{1}{2}A_+,,[A_+,R_+]=0,,[A_+,R_{}]=D_{}$$ (43) along with the Hermitian conjugates of the commutators in Eqs. (41), (42), and (43) close the algebra: the ten independent generators; $`D_+`$, $`D_{}`$, $`L_+`$, $`L_{}`$, $`L_0`$, $`R_+`$, $`R_{}`$, $`R_0`$, $`A_+`$, and $`A_{}`$ form a Lie algebra, the symplectic algebra $`Sp(4)`$. In terms of these generators the most general mass term of the Hamiltonian can be written as $$H_m=m_D(D_++D_{})+m_L(L_++L_{})+m_R(R_++R_{}).$$ (44) In writing Eq.(44) we ignored the phases of the masses. These can easily be incorporated as e.g. $`m_L(e^{i\chi }L_++L_{}e^{i\chi })`$. One can write down a number of 4-dimensional matrix realizations of the Sp(4) algebra. Here we present two of them. The first one can be expressed using the $`2\times 2`$ Pauli matrices as : $$D_0=\frac{1}{2}\left(\begin{array}{cc}\sigma _3& 0\\ & \\ 0& \sigma _3\end{array}\right),D_+=\frac{1}{2}\left(\begin{array}{cc}0& \sigma _+\\ & \\ \sigma _+& 0\end{array}\right),D_{}=\frac{1}{2}\left(\begin{array}{cc}0& \sigma _{}\\ & \\ \sigma _{}& 0\end{array}\right),$$ (45) $$\stackrel{}{L}=\frac{1}{2}\left(\begin{array}{cc}0& 0\\ & \\ 0& \stackrel{}{\sigma }\end{array}\right),\stackrel{}{R}=\frac{1}{2}\left(\begin{array}{cc}\stackrel{}{\sigma }& 0\\ & \\ 0& 0\end{array}\right),$$ (46) $$A_+=\frac{1}{2}\left(\begin{array}{cc}0& \sigma _3+1\\ & \\ \sigma _31& 0\end{array}\right),A_{}=\frac{1}{2}\left(\begin{array}{cc}0& \sigma _31\\ & \\ \sigma _3+1& 0\end{array}\right).$$ (47) In these expressions $`1`$ is the $`2\times 2`$ unit matrix. The second matrix realization is given as $$D_0=\frac{1}{2}\left(\begin{array}{cc}1& 0\\ & \\ 0& 1\end{array}\right),D_+=\left(\begin{array}{cc}0& \sigma _1\\ & \\ 0& 0\end{array}\right),D_{}=\left(\begin{array}{cc}0& 0\\ & \\ \sigma _1& 0\end{array}\right);$$ (48) $$L_0=\frac{1}{4}\left(\begin{array}{cc}\sigma _3+1& 0\\ & \\ 0& \sigma _31\end{array}\right),L_+=\frac{1}{2}\left(\begin{array}{cc}0& \sigma _3+1\\ & \\ 0& 0\end{array}\right),L_{}=\frac{1}{2}\left(\begin{array}{cc}0& 0\\ & \\ \sigma _3+1& 0\end{array}\right),$$ (49) $$R_0=\frac{1}{4}\left(\begin{array}{cc}\sigma _3+1& 0\\ & \\ 0& \sigma _31\end{array}\right),R_+=\frac{1}{2}\left(\begin{array}{cc}0& \sigma _3+1\\ & \\ 0& 0\end{array}\right),R_{}=\frac{1}{2}\left(\begin{array}{cc}0& 0\\ & \\ \sigma _3+1& 0\end{array}\right),$$ (50) $$A_+=\frac{1}{2}\left(\begin{array}{cc}\sigma _{}& 0\\ & \\ 0& \sigma _+\end{array}\right),A_{}=\frac{1}{2}\left(\begin{array}{cc}\sigma _+& 0\\ & \\ 0& \sigma _{}\end{array}\right).$$ (51) Since there is only one four-dimensional representation of $`Sp(4)`$ these two matrix realizations can be transformed into one another by the unitary transformation $$U=\left(\begin{array}{cccc}0& 0\hfill & 0& \hfill 1\\ & & & \\ 0& 1\hfill & 0& \hfill 0\\ & & & \\ 0& 0\hfill & 1& \hfill 0\\ & & & \\ 1& 0\hfill & 0& \hfill 0\end{array}\right).$$ (52) Using the representation given in Eqs. (48), (49), (50), and (51) one can also write $`H_m`$ of Eq. (44) in matrix form $$H_m=\left(\begin{array}{cccc}0& 0\hfill & m_L& \hfill m_D\\ & & & \\ 0& 0\hfill & m_D& \hfill m_R\\ & & & \\ m_L& m_D\hfill & 0& \hfill 0\\ & & & \\ m_D& m_R\hfill & 0& \hfill 0\end{array}\right).$$ (53) Most discussions of the neutrino mass matrix start with Eq. (53) and its generalization to many flavors (see, for example Ref. . A careful discussion which pays particular attention to the phases is given in Ref. ). $`Sp(4)`$ is isomorphic to $`Spin(5)`$ which is the universal covering group of $`SO(5)`$. Representations of $`Spin(5)`$ can be considered as spinor representations of $`SO(5)`$. It may be interesting to point out that these spinor representations can be realized using quaternions or, equivalently, using Pauli matrices. The four dimensional fundamental representation of $`Sp(4)`$ is explicitly given by Eqs. (48) through (51). To exhibit the $`SU(2)`$-doublet structure of this representation we rewrite these equations as $$D_\pm =\frac{1}{2}\sigma _1\sigma _\pm ,$$ (54) $$\stackrel{}{L}=\frac{1}{4}(1+\sigma _3)\stackrel{}{\sigma },$$ (55) $$\stackrel{}{R}=\frac{1}{4}(1\sigma _3)\stackrel{}{\sigma },$$ (56) and $$A_\pm =\left[\sigma _{}\frac{1}{4}(1+\sigma _3)\right]\left[\sigma _\pm \frac{1}{4}(\sigma _31)\right].$$ (57) In these equations the notation $`A\sigma _1`$, etc. stands for the $`4\times 4`$ block matrix where the elements of $`\sigma _1`$ are replaced by the $`2\times 2`$ matrix $`A`$. Full details and the representation theory of spinor representations is covered in standard texts (see for example ). ## III The Pauli-Gürsey transformation and the see-saw mechanism In 1957, W. Pauli introduced a 3-parameter group which relates a Dirac spinor to its charge-conjugation. Later F. Gürsey showed that this group can be extended to particles with mass and charge. The Pauli-Gürsey transformation is given by the 3-parameter transformation $$\psi \psi ^{}=a\psi +b\gamma _5\psi ^c$$ (58) where the parameters $`a`$ and $`b`$ satisfy the unitarity condition $$|a|^2+|b|^2=1.$$ (59) In this section we elaborate on the relationship between the Pauli-Gürsey transformation and the neutrino mass matrix. As the discussion in the previous section indicates the physical meaning of the $`SU(2)`$ algebras generated by the $`D`$’s, $`L`$’s, and $`R`$’s is clear. They generate the neutrino mass matrix. To illustrate the role of the $`SU(2)`$ algebra spanned by $`A_\pm `$ and $`A_0`$ let us consider a general element of the associated $`SU(2)`$ group $$\widehat{U}=e^{\tau ^{}A_{}}e^{log(1+|\tau |^2)A_0}e^{\tau A_+}e^{i\phi A_0},$$ (60) where $`\tau `$ and $`\phi `$ are arbitrary complex and real numbers respectively. Under this $`SU(2)`$ rotation the fermion field transforms into $$\psi \psi ^{}=\widehat{U}\psi \widehat{U}^{}=\frac{e^{i\phi /2}}{\sqrt{1+|\tau |^2}}[\psi \tau ^{}\gamma _5\psi ^c].$$ (61) This is a Pauli-Gürsey transformation with the parameters $`a={\displaystyle \frac{e^{i\phi /2}}{\sqrt{1+|\tau |^2}}},b={\displaystyle \frac{\tau ^{}e^{i\phi /2}}{\sqrt{1+|\tau |^2}}}.`$ (62) Note that if the mass term is considered a pairing interaction the Pauli-Gürsey transformation can be viewed as the associated Bogoliubov transformation. In light of Eqs. (61) and (62) from now on we designate the $`SU(2)`$ algebra spanned by $`A_\pm `$ and $`A_0`$ as $`SU(2)_{\mathrm{PG}}`$. It is also easy to show that the $`U(1)`$ algebra spanned by $`D_0`$ commutes with $`SU(2)_{\mathrm{PG}}`$. At this point the role of various components of the $`Sp(4)`$ algebra is identified. It is easy to show that $$D_0=\frac{1}{2}d^3x(\psi ^{}\gamma _5\psi )$$ (63) Hence the $`U(1)_\chi `$ algebra spanned by $`D_0`$ is a chirality transformation, the $`SU(2)_{\mathrm{PG}}`$ generates the Pauli-Gürsey transformation and the most general neutrino mass Hamiltonian sits in the $`Sp(4)/SU(2)_{\mathrm{PG}}\times U(1)_\chi `$ coset. It turns out that the $`SU(2)_{\mathrm{PG}}`$ has another interesting function, namely that it generates the see-saw mechanism. To illustrate this we first rewrite Eq. (44) in the form $`H_m=m_D(D_++D_{})`$ $`+`$ $`\left({\displaystyle \frac{m_L+m_R}{2}}\right)\left[(L_++L_{})+(R_++R_{})\right]`$ (64) $`+`$ $`\left({\displaystyle \frac{m_Lm_R}{2}}\right)\left[(L_++L_{})(R_++R_{})\right].`$ (65) One can easily show that under an $`SU(2)_{\mathrm{PG}}`$ rotation one gets $$L_+L_+^{}=\frac{e^{i\phi }}{1+|\tau |^2}\left[L_++\tau D_++\tau ^2R_+\right];$$ (66) $$R_+R_+^{}=\frac{e^{i\phi }}{1+|\tau |^2}\left[R_+\tau ^{}D_++(\tau ^{})^2L_+\right];$$ (67) $$D_+D_+^{}=\frac{1}{1+|\tau |^2}\left[(1|\tau |^2)D_++2\tau R_+2\tau ^{}L_+\right].$$ (68) From Eqs. (66) and (67) it follows that under an $`SU(2)_{\mathrm{PG}}`$ rotation one has $$\left[(L_++L_{})+(R_++R_{})\right]\left[(L_+^{}+L_{}^{})+(R_+^{}+R_{}^{})\right]=\left[(L_++L_{})+(R_++R_{})\right]$$ (69) and $`[(L_++L_{})`$ $``$ $`(R_++R_{})][(L_+^{}+L_{}^{})(R_+^{}+R_{}^{})]`$ (70) $`=`$ $`{\displaystyle \frac{1}{1+\tau ^2}}\left[(1\tau ^2)\left((L_++L_{})(R_++R_{})\right)+2\tau \left(D_++D_{}\right)\right].`$ (71) In writing down Eqs. (69) and (70) we took $`\phi =0`$ and $`\tau `$ to be real assuming the masses in Eq. (64) have no phases. If phases are included one should restore $`\phi `$ and a complex $`\tau `$. Under the $`SU(2)_{\mathrm{PG}}`$ rotation the mass Hamiltonian of Eq. (64) transforms into $`H_mH_m^{}`$ $`=`$ $`\widehat{U}H_m\widehat{U}^{}=m_D(D_+^{}+D_{}^{})+\left({\displaystyle \frac{m_L+m_R}{2}}\right)\left[(L_+^{}+L_{}^{})+(R_+^{}+R_{}^{})\right]`$ (72) $`+`$ $`\left({\displaystyle \frac{m_Lm_R}{2}}\right)\left[(L_+^{}+L_{}^{})(R_+^{}+R_{}^{})\right]`$ (73) $`=`$ $`\left({\displaystyle \frac{m_L+m_R}{2}}\right)\left[(L_++L_{})+(R_++R_{})\right]`$ (74) $`+`$ $`{\displaystyle \frac{1}{1+\tau ^2}}\left[m_D(1\tau ^2)+\left(m_Lm_R\right)\tau \right](D_++D_{})`$ (75) $`+`$ $`{\displaystyle \frac{1}{1+\tau ^2}}\left[2m_D\tau +\left({\displaystyle \frac{m_Lm_R}{2}}\right)(1\tau ^2)\right]\left[(L_++L_{})(R_++R_{})\right]`$ (76) One can eliminate various terms from Eq. (72) by a suitable choice of $`\tau `$. For example, if one would like to eliminate the Dirac mass term $`(D_++D_{})`$ one can choose $`\tau `$ to make its coefficient zero. Introducing the angle $`\delta `$ $$\tau =\mathrm{tan}\delta ,$$ (77) one finds that the choice $$\mathrm{tan}2\delta =\frac{2m_D}{(m_Lm_R)},,$$ (78) achieves this goal. The choice in Eq. (78) corresponds to: $$\mathrm{cos}2\delta =\frac{(m_Lm_R)}{[4m_D^2+(m_Lm_R)^2]^{1/2}},\mathrm{sin}2\delta =\frac{2m_D}{[4m_D^2+(m_Lm_R)^2]^{1/2}},$$ (79) Note that this solution exists even in the limiting case of $`m_L=m_R`$. Substituting Eq. (79) into Eq. (72) one obtains $`H_mH_m^{}`$ $`=`$ $`\widehat{U}H_m\widehat{U}^{}=\left({\displaystyle \frac{m_L+m_R}{2}}\right)\left[(L_++L_{})+(R_++R_{})\right]`$ (80) $``$ $`{\displaystyle \frac{1}{2}}\left[4m_D^2+\left(m_Lm_R\right)^2\right]^{1/2}\left[(L_++L_{})(R_++R_{})\right].`$ (81) Making the choice $`m_L=0`$ and $`m_Rm_D`$ one obtains the standard see-saw result $$H_m^{}m_R(R_++R_{})\frac{m_D^2}{m_R}(L_++L_{}).$$ (82) So far the discussion was for a single neutrino flavor. For three neutrino flavors one can introduce three commuting copies of the $`Sp(4)`$ algebra and the similar arguments follow. It is possible to introduce a single $`Sp(4)`$ algebra for three flavors if the individual masses of the mass eigenstates are equal up to phases. This case seems to be too restrictive for model building as it is at variance with the recent observations. ## IV Conclusions We showed that the neutrino mass problem can be formulated algebraically. The algebra is Sp(4). A subgroup of this algebra generates the see-saw mechanism and the neutrino mass Hamiltonian sits in the leftover coset space. We should emphasize that even though in writing down the symplectic algebra we utilized a framework which is symmetric under the exchange of the left- and right-handed fields, the algebraic basis we introduced does not specify the values of $`m_L`$ and $`m_R`$. In particular the value $`m_L=0`$ is not excluded. Note that in the special case of $`m_L=m_R`$ and $`m_D=0`$, Eq. (69) indicates that a general Pauli-Gürsey transformation given in Eq. (60) leaves the mass-term invariant. Indeed in this limit only the diagonal $`SU(2)_{L+R}`$ subgroup of $`SU(2)_L\times SU(2)_R`$ (spanned by the $`L`$’s and $`R`$’s respectively) is sufficient to describe the mass Hamiltonian. This $`SU(2)_{L+R}`$ commutes with $`SU(2)_{PG}`$. Our analysis here concerns the neutrino mass, not the mixing matrix. Recent observations of the atmospheric neutrinos very strongly suggest that muon neutrino maximally mixes with another neutrino. In this case one can show that one of the mass eigenstates decouples from the problem reducing the dimension of flavor space by one , which suggests the existence of an underlying symmetry. It would be interesting to see if the $`Sp(4)`$ or a higher symmetry could make a statement about the mixing matrix. Work along this direction is currently in progress. ## ACKNOWLEDGMENTS This work was supported in part by the U.S. National Science Foundation Grant No. PHY-9605140 at the University of Wisconsin, and in part by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation. A.B.B. acknowledges the support of the Alexander von Humboldt-Stiftung. The work of N.Ö was in part supported by the Scientific and Technical Research Council of Turkey (TUBITAK) under the project No. TBAG-1247. A.B.B. thanks to the Max-Planck-Institut für Kernphysik for the very kind hospitality.
warning/0003/astro-ph0003048.html
ar5iv
text
# WR 146 - Observing the OB-type companion ## 1 Introduction Wolf-Rayet stars are surrounded by dense stellar winds giving rise to free-free emission extending from IR to radio wavelengths. Typically, this emission is characterised by a power-law spectrum of the form $`S_\nu \nu ^\alpha `$, with values of the spectral index $`\alpha +0.7+0.8`$, and radio brightness temperatures $`10^4`$ K. A small number of WR stars have radio emission that exhibits quite different properties: negative spectral indices and brightness temperatures $`10^6`$ K or higher, properties that are characteristic of non-thermal emission. WR 146 is a member of this group, which includes WR 125, WR 140 and WR 147. The radio emission from WR 146 was first resolved in high resolution observations with MERLIN (Dougherty et al. 1996, hereafter Paper I). These 5-GHz observations revealed two components, N<sub>5</sub> and S<sub>5</sub>, separated by $`120`$ milli-arcseconds (mas). The flux of S<sub>5</sub> was consistent with that estimated from extrapolation of the IR-millimetre spectrum arising from the free-free emitting envelope around the WR star. The brightness temperature of N<sub>5</sub> ($`10^6`$K) identified the nature of the emission from this component as non-thermal. An optical spectrum showed evidence for absorption lines at H$`\delta `$ and H$`\gamma `$, which we attributed to an early-type companion to the WR star. This led us to hypothesize in Paper I that the non-thermal emission arose from a population of relativistic electrons, accelerated in a wind-wind collision region where the wind of the WR star and the companion interacted (e.g. Eichler & Usov 1993). To be consistent with such a model, we suggested the companion lay at the same position angle as N<sub>5</sub> from S<sub>5</sub>, but slightly further away from the WR star. The presence of a companion was confirmed in optical imaging with the Hubble Space Telescope (HST) by Niemela et al. (1998). They observed two stars, WR146A and B (hereafter S<sub>O</sub> and N<sub>O</sub> respectively), at the same position angle as the radio sources but separated by $`168`$ mas. Under the assumption that the southern sources in both the HST and MERLIN images are coincident, these observations place the non-thermal source between the two stellar images, strongly supporting wind-wind collision as the origin of the non-thermal emission. Its position relative to the two stellar components ($`120`$ mas from S<sub>O</sub> and $``$ 48 mas from N<sub>O</sub>) is where the dynamical pressure of the two stellar winds is balanced. This indicates that the momentum of S<sub>O</sub>’s stellar wind is $``$ 0.1 times that of N<sub>O</sub>. With the wind velocity of the WC star in WR 146 ($``$ 2900 km s<sup>-1</sup>, Eenens & Williams 1994) being greater than that of a typical OB star, and the expectation that the mass-loss rate of a WR star would be greater than that of an OB star, this strongly supports the identification of S<sub>O</sub> with the WC6 star in Paper I and identifies N<sub>O</sub> with an OB companion having a lesser wind momentum. The momentum ratio lead Niemela et al. to infer that the wind momentum of the companion was more appropriate to a star of early O or Of type than a late O-type main-sequence star. Taking account of the photometry, they suggested an O6-O5 V-III spectral type for the companion. The photometry by Niemela et al. showed the two stars to be equally bright in B but that S<sub>O</sub> was redder than N<sub>O</sub> in (B–V) and (U–B). The (near-zero) magnitude difference between the two blue images falls between two very different estimates of the WR:O light ratio deduced from spectra of WR 146 which included both stars. On the one hand, in Paper I, we measured equivalent widths of H$`\delta `$ and H$`\gamma `$ in our blue spectrum of WR 146 to be $`W_\lambda `$0.9Å. Comparison of these with those ($`2.5`$Å) typical of mid-to-late O stars in the Walborn & Fitzpatrick (1990) atlas indicated that the O-star spectrum was diluted and suggested that the continuum of the WC star was twice as bright as that of the O companion in the blue. On the other hand, Willis et al. (1997) used a spectrum of lower resolution but longer wavelength coverage than presented in Paper I to determine the WR:O light ratio from comparison of the equivalent widths of the emission lines with those of two single WC6 stars. They found that the O star was brighter than the WC star, reporting a continuum light ratio (WR:O) of 1:(2$`\pm `$1). The discrepancy in the light ratios of the WC6 and O components derived from the dilution of O star absorption lines and WC6 star emission lines is significant and probably too great to arise purely from the observational uncertainties. Leaving aside the question of whether one of the stars varied between the two observations, the principal uncertainties in two spectroscopic light-ratio determinations come from whether the intrinsic line strengths of the components are indeed equal to those of the comparisons adopted. The discrepancy could be reduced if the equivalent widths of H$`\delta `$ and H$`\gamma `$ in the O star were weaker than the 2.5Å adopted or if the emission lines in the WC6 component were weaker than those of other WC6 stars. Both possibilities are plausible. The strengths of the emission lines in WR 146 may well be atypical for its WC6 type: both Eenens & Williams (1992) and Willis et al. found anomalously low C/He abundances for this star. On the other hand, the strengths of H$`\gamma `$ and H$`\delta `$ in O-type stars do depend on spectral type and luminosity class, which are not known directly for the companion in WR 146. We note that the “O8.5V” adopted by Willis et al. comes not from the star’s spectral lines but from its luminosity inferred from that of the WC6 star and the continuum light ratio. We therefore re-observed the blue spectrum of WR 146 with the William Herschel Telescope in an attempt to determine the spectral type and, hopefully, luminosity of the early-type companion to the WR star. These data, together with the photometry by Niemela et al., will give us a better idea of the intrinsic properties of the companion for comparison with the radio observations. We also re-observed WR 146 at radio wavelengths, extending the frequency coverage of Paper I, to search for the wind flux from the companion, and study the characteristics of the two previously observed radio components with the aim of furthering our understanding of the colliding wind phenomenon. ## 2 Radio Wavelength Study ### 2.1 The observations Multi-frequency observations of WR 146 at 1.4, 5, 8.4, and 22 GHz were taken on 1996 October 26 using the A-configuration of the NRAO Very Large Array (VLA). Observations of the nearby radio-bright quasar 2005+403 were interleaved with the observations of the target source for phase-referenced calibration of the antennae gains. The absolute flux scale was determined by observation of 3C48, assumed to have fluxes of 15.970, 5.516, 3.226 and 1.174 Jy at 1.4, 5, 8.4, and 22 GHz respectively. In addition, archival data from MERLIN at 5 GHz have been analysed. A 4-hour on-source observation at 5 GHz was obtained with MERLIN on 1992 December 26, along with frequent observations of the phase calibration source 2005+403. Following the standard procedure for calibrating MERLIN 5-GHz data, the flux scale was established by observation of 3C286 with an assumed total flux of 7.309Jy (7.020 Jy on the MK2-Tabley baseline) and the unresolved source 0552+398, which had a derived “bootstrap” flux of 5.786 Jy. The bootstrap fluxes for 2005+403 are given in Table LABEL:tab:fluxes. It is possible to check the accuracy of the derived flux scales in the 5- and 8.4-GHz observations. Comparison of the “bootstrap” fluxes of 2005+403 at these frequencies to those given in the University of Michigan Radio Astronomy Observatory (UMRAO) flux database shows that, where closely contemporaneous observations are available, the agreement with UMRAO fluxes at these frequencies is within $`1\%`$. Given the excellent agreement at these two frequencies, it is reasonable to assume that the flux scales at the other observed frequencies are good. Aside from the initial amplitude calibration of the MERLIN observations, the calibration and subsequent imaging of the data were carried out using the NRAO AIPS software package. The complex antennae gains were initially derived for 2005+403 and then interpolated (phase-referenced) to the observations of WR 146. In addition, several iterations of phase-only self-calibration were used to refine the antennae gains during the MERLIN observations of WR 146 to improve the dynamic range of the final synthesized images. At 1.5 and 4.9 GHz the VLA data reveal an unresolved source, whereas at 8.4 GHz the radio emission is marginally resolved. However, WR 146 is well resolved into a double radio source at 22 GHz. The final synthesized 22-GHz image is shown in Fig. 2. We identify the northern and southern components as N<sub>22</sub> and S<sub>22</sub>. The visibility data suggest both these components are resolved (Figure 3). Gaussian model fits to the visibilities give the diameter of the radio emitting regions as $`48\pm 11`$ and $`78\pm 19`$ mas respectively (see Paper I for a description of the method). The large relative uncertainty in these sizes is a direct consequence of the high rms uncertainty per visibility at 22 GHz. The fluxes from the VLA observations were determined by Gaussian source fitting using the AIPS routine JMFIT. The measured fluxes are given in Table LABEL:tab:fluxes. The final synthesized MERLIN image at 5 GHz is shown in Fig 1. The image reveals two components of emission, identified as N<sub>5</sub> and S<sub>5</sub>. In the MERLIN observation, the flux and size of component N<sub>5</sub> were determined by Gaussian model fitting of the visibility data. The derived diameter for N<sub>5</sub> is $`42\pm 1`$ mas. Gaussian fitting to the image data were used for the source parameters of S given in Table 1. Qualitatively, the 1992 MERLIN image is consistent with the 5-GHz observation from 1995 April 29 (Paper I) though there is some evidence that N<sub>5</sub> may have decreased in flux by $`3`$ mJy between 1992 and 1995. The flux of S<sub>5</sub> appears to have decreased between the two epochs. However, the difference is only $`2\sigma `$, much less than the $`5\sigma `$ threshold typically adopted as evidence for variation. Therefore, we adopt the mean value of $`2.0\pm 0.2`$ mJy for the flux of S<sub>5</sub>. Variations in the total 5-GHz emission from WR 146 have been observed using the WSRT over the last decade that are attributed to variations in the non-thermal component (Setia-Gunawan et al. 2000). Both our MERLIN 5-GHz total flux values are consistent with the WSRT observations. Variability could account for the slightly higher 5-GHz flux at the epoch of the VLA observations. The positions of the components observed at 5 GHz and 22 GHz are quoted in Table LABEL:tab:positions. The absolute position of the northern component in the various observations was deduced from the phase-reference only calibrated images. Absolute position information is lost during the self-calibration process, though the relative position is preserved. These were deduced from the images shown in Fig. 1 and Fig. 2. ### 2.2 Identifying the radio components The nominal resolution of A-configuration at 22-GHz observations is very similar to that of MERLIN at 5 GHz, allowing direct comparison of these data. At 22 GHz, the VLA observations reveal two components N<sub>22</sub> and S<sub>22</sub>, very similar in appearance to the MERLIN 5-GHz data. However, the relative positions of the two 22-GHz components show them to be significantly further apart than those from the MERLIN 5-GHz data, leading to the question of whether the 22-GHz sources are the same as those observed at 5 GHz. The brightness temperatures of N<sub>22</sub>, S<sub>22</sub> and S<sub>5</sub> are $`1.6\times 10^4`$ K, $`5000`$ K and $`4000`$ K respectively, all consistent with an origin in a photo-ionized circumstellar envelope where the photo-ionization equilibrium temperature is typically $`10^4`$ K. In contrast, this is about two orders of magnitude lower than the brightness temperature of $`1.3\times 10^6`$ K for N<sub>5</sub>, which clearly indicates a non-thermal origin. The separation of S<sub>22</sub> and N<sub>22</sub> ($`162\pm 8`$ at a position angle of $`22\pm 4^{}`$) is very close to that ($`168\pm 31`$ mas at position angle $`21\pm 4^{}`$) of the optical components (N<sub>O</sub> and S<sub>O</sub>) observed with the HST. There is no evidence of intrinsic proper motion in the source at 22 GHz in recent (1999) observations of the source (A. Fink, private communication). The coincidence of the optical and 22-GHz positions, the lack of evidence of intrinsic proper motion, and the thermal nature of the emission from the two 22-GHz components leads us to conclude that the 22-GHz components are the stellar winds associated with the two stellar components imaged with the HST by Niemela et al. (1988). This is the first time that the stellar wind of the companion in a WR binary has been spatially resolved. In Paper I we concluded that S<sub>5</sub> was the thermal emission from the WR-star wind. Indeed, if we take the mean flux value for S<sub>5</sub> from our two epochs of MERLIN observations, the spectral index between 22 and 5 GHz for the southern component is $`+0.82\pm 0.14`$, consistent with those observed in the stellar winds of other WC-type WR stars (e.g. Williams 1996). We estimate the diameter of S<sub>5</sub> to be $`190`$ mas from the diameter of the S<sub>22</sub>, since angular size $`\nu ^{0.6}`$ for $`\alpha 0.8`$. This is approximately four times larger than that given in Paper I. However, we feel the diameter presented here is a more reliable estimate. Additionally, in Paper 1 we identified N<sub>5</sub> as non-thermal emission from a wind-wind collision region. If we overlay the 22-GHz and 5-GHz data assuming two southern components originate in the WR star wind (see Figure 4) we can see that the relative separation of the components supports such a model, where the wind-collision region must fall between the two stars. ### 2.3 Wind parameters for both stellar components For a steady state, smooth, fully ionized stellar wind having a radial $`r^2`$ ion density distribution, the free-free radio flux $`S_\nu `$ is related to the stellar-wind density i.e. $`\dot{M}/v`$ by $$S_\nu =2.32\times 10^4(\gamma g_{\nu ,T_e}\nu )^{\frac{2}{3}}\left(\frac{\dot{M}Z}{\mu v}\right)^{\frac{4}{3}}d^2\mathrm{mJy},$$ (1) where $`\nu `$ is the frequency in Hz, $`d`$ is the distance in kpc, $`\dot{M}`$ is the mass-loss rate in M y<sup>-1</sup>, $`v`$ the terminal velocity of the wind in km s<sup>-1</sup>; $`\gamma `$, $`Z`$, $`\mu `$ and $`g_{\nu ,T_e}`$ are respectively, the number of electrons per ion, the mean charge per ion, the mean atomic weight and the Gaunt free-free factor, a function of electron temperature and frequency. It follows from (1) that $$\frac{\dot{M}}{v}\left(\frac{\mu ^2}{Z^2g_{\nu ,T_e}\gamma }\right)^{\frac{1}{2}}S^{\frac{3}{4}}.$$ (2) Thus, the measurement of radio emission from the two stellar wind components in the WR 146 system leads to the density ratio of the two winds. If the stellar winds are clumped rather than smooth, we can allow for this by replacing $`\dot{M}`$ in the above with $`(\dot{M}/\sqrt{f})`$, where $`f`$ is the wind filling factor, assumed to be constant over the radio-emitting region. In the case of the collision of two spherical winds at terminal velocity, the contact discontinuity between the two winds intersects the line of centres between the stars at the point of momentum balance. If the projected separation of the WR and OB star is $`D\mathrm{cos}i`$, and that of the OB companion from the non-thermal region is $`r_{\mathrm{OB}}\mathrm{cos}i`$, we can then write $$r_{\mathrm{OB}}/D=\frac{\eta ^{\frac{1}{2}}}{1+\eta ^{\frac{1}{2}}},$$ (3) where $`\eta `$ is the ratio of WR and OB-companion wind momenta, $`(\dot{M}v)_{\mathrm{OB}}/(\dot{M}v)_{\mathrm{WR}}`$. The wind-collision geometry and equation 3 would also be affected by clumping, but this is beyond the scope of the present study. By defining the ratio of the OB star and WR-star wind densities, $`(\dot{M}/v)_{\mathrm{OB}}/(\dot{M}/v)_{\mathrm{WR}}`$, as $`\xi `$, then $$\frac{\dot{M}_{\mathrm{OB}}}{\dot{M}_{\mathrm{WR}}}=(\eta \xi )^{\frac{1}{2}}\&\frac{v_{\mathrm{OB}}}{v_{\mathrm{WR}}}=(\frac{\eta }{\xi })^{\frac{1}{2}}.$$ (4) Knowing $`\dot{M}_{\mathrm{WR}}`$ and $`v_{\mathrm{WR}}`$ we can now uniquely determine $`\dot{M}_{\mathrm{OB}}`$ and $`v_{\mathrm{OB}}`$ that satisfy simultaneously both $`\xi `$ and $`\eta `$. An attractive property of these ratios is that they are independent of distance, typically a very uncertain parameter for WR stars and particularly for WR 146. For the OB companion we will assume that hydrogen and helium are, respectively, singly and doubly ionized, leading to values $`\gamma =1.1`$, $`Z=1.15`$, $`\mu =1.34`$ (Lamers & Leitherer 1993) and $`T_e10^4`$ K. Taking the values of $`\gamma =1.15,Z=1.2`$, $`\mu =5.29`$, $`T_e=\mathrm{8\hspace{0.17em}000}`$ K from Willis et al. (1997) for the WR star, and that the fluxes of N<sub>22</sub> and S<sub>22</sub> are those of the OB star and the WC star respectively, implies that $`\xi =0.36\pm 0.19`$. From Table 2, $`D`$ is $`162\pm 8`$ mas and $`r_{OB}`$ is $`40\pm 9`$ mas, so equation 3 gives $`\eta =0.11\pm 0.03`$. Thus, the ratios of mass-loss rates and wind velocities are $`0.20\pm 0.06`$ and $`0.56\pm 0.17`$ respectively. The largest uncertainty in these ratios arises from $`\xi `$. Though this requires knowledge of the relative metallicity and ionization structure within the two winds, the values of $`Z`$ and $`\gamma `$ are closely the same and effectively cancel. On the other hand, the mean molecular weights of the two winds are quite different and contributes a factor $`2`$ to the ratio. The relative uncertainty in the observed fluxes provides the bulk of the uncertainty in $`\xi `$. From the deduced velocity ratio and the terminal wind velocity measured for the WC6 star ($`2900`$ km s<sup>-1</sup>, Eenens & Williams 1994), we derive a terminal velocity of the OB wind of $`1600\pm 480`$ km s<sup>-1</sup>. This is consistent with those determined for late O-type stars of any luminosity class (Prinja, Barlow & Howarth 1990). However, the deduced ratio (0.19) of mass-loss rates implies a very high mass-loss rate for the OB star, irrespective of the distance to WR 146. For example, if the WC6 star has a mass-loss rate comparable to the average rate derived from radio or from “Standard Model” analyses of optical emission lines ($`4\times 10^5`$ M y<sup>-1</sup>, Willis 1999), that of the OB star would be $`8\times 10^6`$ My<sup>-1</sup>. This is comparable to that determined for the O8f supergiant HD 151804 (Bieging, Abbott & Churchwell 1989, Crowther & Bohannan 1997) and almost $`50\times `$ greater than the mass-loss rates typical of main-sequence O8 stars ($`1.7\times 10^7`$ M y<sup>-1</sup>, Howarth & Prinja 1989). This result still holds if the true mass-loss rates of WR stars are lower than the value cited owing to clumping since it is really the ratio of $`(\dot{M}/\sqrt{f})`$s that has been determined. Only if the OB-star wind was clumped would its mass-loss rate be reduced by a factor of $`\sqrt{f}`$ . Throughout the paper we have identified the southern and northern sources as the WR and OB star respectively. If these identifications were switched i.e. WR star is to the north, and the OB star to the south, then the deduced wind velocity of the OB star is greater than $`\mathrm{18\hspace{0.17em}000}`$ km s<sup>-1</sup> and the mass-loss rate is 1.4 times that of the WR star. Clearly these parameters are unacceptable given our knowledge of OB-star winds and we dismiss such a source identification as untenable. The mass-loss rates determined from the measured free-free wind fluxes depend heavily on the adopted distance to WR 146. Therefore, we first re-examine the luminosity of the system using new optical observations. ## 3 Optical Spectroscopy ### 3.1 The observations The spectrum of WR 146 was taken with the blue arm of the ISIS spectrograph on the William Herschel Telescope (WHT) on 1999 June 19. As in Paper I, we observed in the blue, where the relative contribution of the OB companion would be greatest, and avoided the strong $`\lambda `$4650Å emission feature from the WR star. The R1200B grating and 1.1 arcsec slit gave a resolution of 0.9 Å. Four integrations, each of 750s, were taken of the star, interleaved with observations of an argon discharge tube for wavelength calibration. The detector used was a $`4200\times 2048`$ EEV42 device. We present the new spectrum of WR 146 in Fig. 5 and list absorption lines in Table 3. The spectrum has been smoothed giving a resolution $``$ 1.0 Å (FWHM), as measured from the widths of the narrow interstellar CH, CH<sup>+</sup> and Ca i lines. The FWHM of the $`\lambda `$3933Å Ca ii line is 1.6 Å, the greater width coming from optical depth effects (cf. Jenniskens & Désert 1994). The well developed interstellar spectrum is consistent with the heavy reddening suffered by WR 146. The broad undulations in the continuum are mostly emission lines from the WC6 star, diluted by the light from the companion. The emission lines in WR 146 are unusually broad for its WC6 subtype (cf. Eenens & Williams 1994, Willis et al. 1997), making it easier to measure the absorption spectrum of the companion. The absorption line widths are typically $``$ 5Å FWHM, comparable for example to those of the O9.5II star $`\delta `$ Ori (cf. Voels et al. 1989), which has a relatively high $`v\mathrm{sin}i`$ (144 km s<sup>-1</sup>, Howarth et al. 1997). There is one absorption feature from the WC6 star; the P-Cygni absorption component of the $`\lambda `$3889Å He i line, blue-shifted by 2630 km s<sup>-1</sup> to 3854.5Å. This line arises from the same metastable level as the $`\lambda `$10830Å line from which Eenens & Williams (1994) derived a terminal wind velocity (v) of 2900 km s<sup>-1</sup> by profile fitting and, like the $`\lambda `$10830Å line, is expected to form in the outer regions of the WC6 wind. These results are consistent with the v$`{}_{\mathrm{}}{}^{}=2700`$ km s<sup>-1</sup> measured by Willis et al. from the \[Ne iii\] $`\lambda `$15.5-$`\mu `$m fine-structure line, also formed in the outer wind. The other absorption lines (Table 3) are formed in the OB-companion star. Each of the hydrogen lines is blended with the nearby member of the Pickering series of He ii but the relative weakness of the unblended odd-numbered Pickering line at 4200Å and the closer proximity of the hydrogen-line wavelengths to the observed wavelengths suggests that the He ii contribution is relatively small. Helium i is represented by the $`\lambda \lambda `$ 4472 and 4026Å lines; the latter is blended with He ii and is therefore not very useful for diagnostic purposes. Other He i lines in the Walborn & Fitzpatrick (1990) atlas of hot-star spectra are not seen in our spectrum. The relative strength of the $`\lambda `$ 4070Å C iii triplet is a little surprising but the continuum here is uncertain as it falls between two emission features from the WC6 spectrum. In order to establish the continuum in the regions of each of the absorption lines so that we could measure their equivalent widths, we compared our rectified spectrum with that of another WC6 star, WR 15 (HD 79573), artificially diluted so as to match the spectrum of WR 146. The spectrum of WR 15 was observed with the 1.9m telescope at the South African Astronomical Observatory and will be discussed elsewhere. As a by-product of this matching, we estimated the dilution of the WC6 spectrum in WR 146, finding OB:WR $``$ 3, consistent with the ratio, OB:WR $`=`$2$`\pm `$1, found by Willis et al. (1997) from stronger emission lines between 4660Å and 6560Å. We are chary of putting too much weight on this result as it depends heavily on the rectification process and the presumption that the WC6 emission lines in WR 146 have the same strengths as those in WR 15. Instead, we concentrate on the absorption lines, to be discussed below. ### 3.2 Spectral type and luminosity of the OB companion To investigate the nature of the OB companion, we begin by using the helium-line ratios to estimate its spectral type. Our optical spectrum includes contributions from both N<sub>O</sub> and S<sub>O</sub> but the ratios of the absorption lines in the companion will be equal to those measured from our combined-light spectrum, and independent of the dilution by the WR star, provided that the dilution does not change significantly over the wavelength range of the spectrum. Formal O-star spectra classification is based on the ratio of the $`\lambda `$4541Å He ii and $`\lambda `$4472Å He i lines. Since the former was not available, we used the $`\lambda `$4200Å He ii line from the same series. We formed an empirical calibration of $`\lambda `$4472Å/$`\lambda `$4200Å ratios against spectral type using the line strengths measured for a large ($``$ 100) sample of O stars by Conti & Alschuler (1971) and Conti (1973). Using this calibration, the ratio of our measured equivalent widths (Table 3) indicates a type of O8, with an uncertainty of half a subclass. The type is close to that inferred by Willis et al. from their OB:WR light ratio and an average luminosity for the WC6 star; but this agreement is fortuitous. The luminosity class criterion adopted by Conti & Alschuler for the middle and late-type O stars was the ratio of the Si iv $`\lambda `$4089Å and He i $`\lambda `$4143Å lines. Unfortunately, the silicon line is too weak (W$`{}_{\lambda }{}^{}`$ 0.06Å) to measure with confidence and the $`\lambda `$4143Å line is not seen at all. Another significant He i line that is apparently missing is the $`\lambda `$4388Å singlet line. The strength of this relative to that of the $`\lambda `$4472Å He i line (“singlet to triplet ratio”) has long been known to be sensitive to luminosity owing to the relative overpopulation of the He i 2$`{}_{}{}^{3}P_{}^{o}`$ state in extended atmospheres (e.g. Voels et al. 1989 and references therein). Using the equivalent-width measurements by Conti & Alschuler (1971) and Conti (1973), we examined the ratios, W= W<sub>λ</sub>($`\lambda `$4388)/W<sub>λ</sub>($`\lambda `$4472), formed from their observations of the O7.5, O8 and O8.5 stars in their sample. This ratio was found to be a strong diagnostic for Of stars. All the O7.5–8.5 stars having W$`{}_{}{}^{}<0.25`$ are Of stars, most of them supergiants. Measurement of the $`\lambda `$4472Å line and examination of our spectrum in the region of the $`\lambda `$4388Å line indicates that W $``$ 0.2, strongly suggesting that the O8 star in WR 146 is an Of star, and possibly a supergiant. Unfortunately, the classical Of diagnostic emission features C iii $`\lambda \lambda `$4630–41Å and He ii $`\lambda `$4686Å fall within the strong $`\lambda `$ 4650Å C iii-iv/He ii emission feature from the WC6 star, which extends from 4620Å to 4690 Å in WR 146 (Dessart et al. 2000). This increases the dilution of the OB spectrum to $``$ 1:4 in this critical region and we cannot make a formal Of classification. Further circumstantial support for a high luminosity for the OB star comes from comparison of the $`\lambda `$4472Å He i to H$`\gamma `$ ratio with those derived from the Conti & Alschuler and Conti (1973) datasets. On the other hand, we do not observe emission lines corresponding to the ‘unidentified’ features observed in O6–O9.7 supergiants, including the O8If star HD 151804 (Crowther & Bohannan 1997), at 4486Åand 4504Å. Also, the weakness of the Si iv $`\lambda `$4089Å line is surprising given the other evidence for high luminosity, but not unique: the Si iv line is barely visible in the spectrum of the O7Iaf star Sanduleak 80 (Walborn & Fitzpatrick 1990). Although we cannot assign a luminosity class, the spectrum points to an extended atmosphere and high luminosity, but not as high as that of HD 151804. ### 3.3 Spectral line dilution and the HST photometry With a better idea of the nature of the O8 star, can we reconcile the measured absorption-line strengths in our composite spectrum with the observations by Niemela et al. (1998)? We follow Paper I and Niemela et al. by assigning the northern optical image (N<sub>O</sub>) to the O8 star on the basis of its closer proximity to the non-thermal radio source and the assumption that its stellar wind has a lower momentum than that of the WC6 star. If we assume that the spectrum of S<sub>O</sub> is that of a WC6 star, we can estimate the intrinsic equivalent widths in the O8 spectrum from those observed in our combined-light spectrum and Niemela et al.’s photometry of S<sub>O</sub> and N<sub>O</sub>. We find 1.4Å for the intrinsic equivalent width of H$`\gamma `$, which is closer to the average of those measured by Conti (1973) for O8f stars (1.3$`\pm `$0.3Å) than for ‘non-Of’ O8III–V stars (2.5$`\pm `$0.2Å), another indication of high luminosity properties for the OB star. Our earlier view (Paper I) that the WC6 star was brighter than the OB star was based on the erroneous assumption of intrinsic absorption-line strengths appropriate for a main-sequence OB star. The observation of approximately equal B magnitudes for N<sub>O</sub> and S<sub>O</sub> is not consistent with the light ratios derived from dilution of the emission-line spectrum of the WC6 star by Willis et al. (1997), and supported by our estimate in Section 2.1. This could be resolved also if the emission lines in the WC6 component were 30–40% weaker than those typical of WC6 stars. This is plausible; as already noted, the star appears to have an abnormally low C/He abundance ratio and the dilution analysis by Willis et al. depends heavily on C iii and C iv lines. Therefore, the very different light ratios originally derived from combined-light spectra in Paper I and by Willis et al. can be reconciled with a model in which WR 146 comprises two stars of approximately equal brightness in the blue and relatively weak lines, those of N<sub>O</sub> being consistent with other indicators of its high luminosity. ### 3.4 The nature of the northern component The H$`\gamma `$ strength (together with HST photometry), the He i singlet-to-triplet line ratio, and the very high mass-loss rate all point to an extended atmosphere for N<sub>O</sub>. If it is single, it cannot be a main-sequence star but must be luminous. Alternatively, it may itself be an unresolved binary comprising an O8 star with a third component. If N<sub>O</sub> was a single, luminous star, we can estimate its absolute magnitude from those of O8I/f and O8III stars ($`M_V=6.5`$ and $`5.5`$, Conti 1988; Vacca, Garmany & Shull 1996). For example, combination of an intermediate luminosity, $`M_V=6.0`$, which might be appropriate given the apparent absence of the ‘unidentifed’ 4486Å and 4504Å emission lines, with the photometry by Niemela et al. would give $`M_V=6.2`$ for S<sub>O</sub> and a combined $`M_V=6.9`$ for the WR 146 system. The implied luminosity of the WC6 star, $`M_V=6.2`$ or $`M_v6.0`$ on the narrow-band $`ubv`$ system used for WR stars, is significantly greater than those of other WC6 stars (e.g. mean $`M_v=4.1\pm 0.4`$, van der Hucht, in preparation), but this may be another manifestation of the anomalous nature of the WC6 star in WR 146. Adopting a reddening of $`E_{BV}=2.7`$, consistent with Paper I and Willis et al. (1997), the combined $`M_V`$ implies a distance modulus $`mM=11.2`$, greater than our distance in Paper I ($`mM=10.4`$) but consistent with membership of the Cyg OB2 Association ($`mM=11.2`$, Massey & Thompson 1991; Torres-Dodgen, Tapia & Carroll 1991). Given the angular distance ($`0.5^{}`$) of WR 146 from the centre of the association, stronger evidence is needed to confirm membership. At this distance (1.7 kpc), the mass-loss rates derived for the WC6 and O8 stars are very high: $`1.3\times 10^4`$ My<sup>-1</sup> and $`2.5\times 10^5`$ My<sup>-1</sup> respectively. The latter is three times that of HD 151804, whereas the luminosity of N<sub>O</sub> appears to be lower. Such a system would have stronger emission lines that those in HD 151804, contrary to observations (e.g. 4486Å and 4504Å or H$`\alpha `$ in Dessart et al. 2000, fig 4 compared with Crowther & Bohannan 1997), so the single high-luminosity model for N<sub>O</sub> is problematic. If we consider N<sub>O</sub> to be an unresolved binary in which some of the mass-loss is contributed by a third component, the latter is unlikely to be another O-type star. Since $`\dot{M}L^{1.7}`$ for O stars (Howarth & Prinja 1987), we expect the mass-loss rate of an O+O system to be lower than that of a single star having the same total luminosity. Also, the intrinsic strengths of the H$`\gamma `$ lines from two lower luminosity stars are likely to be greater, in conflict with our combined-light spectrum and the HST photometry. A model wherein the third component is another WC star suffers neither of these problems. Its spectral subtype would have to be similar to that of S<sub>O</sub>, or it would have to be fainter, not to have been detected in the combined-light spectra. Its presence could account for some of the anomalous properties (e.g. low C/He) of the WC6 star reported in the studies referred to. It could also modulate the wind of the O8 star flowing towards the wind interaction region, providing a mechanism for the 3.38-y variability in the non-thermal emission reported by Setia Gunawan et al. (2000). To examine this model, separate spectra of N<sub>O</sub> and S<sub>O</sub>, e.g. with an adaptive optics system, are needed to see whether N<sub>O</sub> has a WC companion (or Of emission lines) and to determine its spectrum without dilution. ## 4 Conclusions From high-resolution observations with the VLA at 22 GHz and with MERLIN at 5 GHz we have observed all three components of the WR 146 system: the OB and WC6 stellar winds and the non-thermal source where they collide. The source geometry and ratio of stellar wind fluxes allow us to determine the ratios of mass-loss rates and wind velocities independent of distance to the system. From these ratios and the observations of the WC6 star, we derive the wind velocity of the OB star to be $`1600\pm 480`$ km s<sup>-1</sup> and its mass-loss rate to be one quarter that of the WC6 star. If the WC6 star has an “average” WR-star mass-loss rate of $`4\times 10^5`$ M y<sup>-1</sup>, that of the OB star would be $`8\times 10^6`$ M y<sup>-1</sup>, suggesting a very high luminosity object. Support for high luminosity comes from the optical spectrum of WR 146, which includes both stars. This shows absorption-line ratios formed in the OB companion suggesting it to be a high-luminosity O8 star. If it is a single star, the inferred luminosity places WR 146 at the distance of the Cyg OB2 association. This gives an anomalously high luminosity for the WC6 star but, given its other anomalies, does not rule out this distance. However, the mass-loss rates determined using this distance ($`1.3\times 10^4`$ My<sup>-1</sup> for the WC6 star and $`2.6\times 10^5`$ My<sup>-1</sup> for the O8 star) are awkwardly high and the latter is probably inconsistent with the spectroscopy. Many, if not all, of the observations could be explained if the companion was itself a binary comprising an O8 star and another WC star. This needs to be tested by separate spectra of the two visual components of WR 146. The presence of an unresolved companion to the O8 star could modulate its wind so as to cause the 3.38-y variability in the non-thermal emission reported by Setia Gunawan et al. (2000). We may not have reached a firm conclusion as to the nature of the stellar companion(s) to the WC star from the new radio and optical observations. However, the ability to study all three radio components of the WR 146 system separately will ensure that this system becomes an archetype for studying the wind-collision phenomenon. ## Acknowledgements This research has made use of data from MERLIN, a UK national facility operated by the University of Manchester, and the William Herschel Telescope, operated on Observatorio del Roque de los Muchachos, La Palma, by the Isaac Newton Group, both on behalf of Particle Physics and Astronomy Research Council; the National Radio Astronomy Observatory Very Large Array, a facility of the National Science Foundation operated under cooperative agreement with Associated Universities, Inc; the University of Michigan Radio Astronomy Observatory supported by the National Science Foundation and the University of Michigan; the South African Astronomical Observatory; and the SIMBAD database, operated at the CDS, Strasbourg, France. The authors are indebted to the referee, Ian Howarth, for his detailed critique of the original manuscript and suggestions for improvements that sharpened our thinking; and to Amy Fink for sharing her observations prior to publication. SMD is indebted to the National Research Council of Canada, Dominion Radio Astrophysical Observatory for supporting this research.
warning/0003/astro-ph0003284.html
ar5iv
text
# A REMARKABLE ANGULAR DISTRIBUTION OF THE INTERMEDIATE SUBCLASS OF THE GAMMA-RAY BURSTS ## ABSTRACT In the article a test is developed, which allows to test the null-hypothesis of the intrinsic randomness in the angular distribution of gamma-ray bursts collected at the Current BATSE Catalog. The method is a modified version of the well-known counts-in-cells test, and fully eliminates the non-uniform sky-exposure function of BATSE instrument. Applying this method to the case of all gamma-ray bursts no intrinsic non-randomness was found. The test also did not find intrinsic non-randomnesses for the short and long gamma-ray bursts, respectively. On the other hand, using the method to the new intermediate subclass of gamma-ray bursts, the null-hypothesis of the intrinsic randomness for 181 intermediate gamma-ray bursts is rejected on the $`96.4\%`$ confidence level. Taking 92 dimmer bursts from this subclass itself, we obtain the surprising result: This ”dim” subclass of the intermediate subclass has an intrinsic non-randomness on the 99.3% confidence level. On the other hand, the 89 ”bright” GRBs show no intrinsic non-randomness. Subject headings: cosmology: observations - gamma rays: bursts ## 1 INTRODUCTION Two results of the last years in the statistics of the gamma-ray bursts (GRBs) are doubtlessly remarkable. The first one concerns the number of subclasses. Recently, two different articles (\[Mukherjee et al. 1998\]; \[Horváth 1998\]) simultaneously suggest that the earlier separation \[Kouveliotou et al. 1993\] of GRBs into short and long subclasses is incomplete. (It is a common practice to call GRBs having $`T_{90}<2`$ s ($`T_{90}>2`$ s) as short (long) GRBs, where $`T_{90}`$ is the time during which 90% of the fluence is accumulated \[Kouveliotou et al. 1993\].) These articles show that, in essence, the earlier long subclass alone should be further separated into a new ”intermediate” subclass ($`2`$ s $`<T_{90}<10`$ s) and into a ”truncated long” subclass ($`T_{90}>10`$ s). (In what follows, the long subclass will contain only the GRBs with $`T_{90}>10`$ s, and the intermediate subclass will be considered as a new subclass.) The second result concerns the angular distribution of GRBs. At the last years several attempts (\[Hartmann et al. 1991\], \[Briggs et al. 1996\]; \[Tegmark et al. 1996b\]; \[Balázs et al. 1998\]; \[Balázs et al. 1999\]) were done either to confirm or to reject the randomness in the angular sky distribution of GRBs being collected at BATSE Catalog (\[Fishman et al. 1994\]; \[Meegan et al. 1998\]). Theoretically, if the intrinsic distribution of GRBs is actually random, an observation of some non-randomness is still expected due to the BATSE non-uniform sky-exposure function (\[Fishman et al. 1994\]; \[Meegan et al. 1998\]). Hartmann et al. (1991), Briggs et al. (1996) and Tegmark et al. (1996b) did not find any statistically significant departure from the randomness. On the other hand, the existence of some non-randomness was confirmed on the $`>99.9\%`$ confidence level by Balázs et al. (1998). This behavior can be caused either purely by instrumental effects or the instrumental effects alone do not explain fully the detected behavior and some intrinsic non-randomnesses should also exist. Balázs et al. (1998, 1999) suggest the second possibility. This conclusion follows from the result that while the short subclass shows a non-randomness, the intermediate + long subclasses do not indicate it. It is difficult to explain such behavior of subclasses by the instrumental effects alone. In this article we will again investigate the angular distribution of GRBs. Trivially, after the discovery of the new intermediate subclass, it is highly required to test the intrinsic randomness in the angular distribution of this new subclass, too. In addition, of course, new different tests, which exactly eliminate the effect of the sky-exposure function, are also required in order to complete the results of Balázs et al. (1998, 1999). The aim of this article is to test the intrinsic randomness in the angular distribution of all GRBs and of the three subclasses separately, too. We will use a modification of the well-known counts-in-cells method. This is a standard and simple statistical test (see, cf., Mészáros (1997) and references therein). The advantage of this method is given by the fact that it allows to eliminate quite simply and exactly the sky-exposure function. The main result of paper will be the surprising conclusion that the intermediate subclass and only this subclass alone suggests a non-randomness on the 96.4% confidence level; its ”dimmer” half even on the 99.3% confidence level. The paper is organized as follows. In Section 2 the method is described. In Section 3 the results of test are presented. Section 4 discusses and summarizes the results of the article. ## 2 THE TEST Assume for the moment that there is no non-uniform sky-exposure function. We separate the sky in declination into $`m_{dec}>1`$ stripes having the same area ($`4\pi /m_{dec}`$ steradian). The boundaries of stripes are the declinations $`\delta _k,k=0,1,\mathrm{},m_{dec}`$, where $`\delta _0=90`$ degree and $`\delta _{m_{dec}}=+90`$ degree, respectively. The remaining values are analytically calculable, and appear symmetrically with respect to $`\delta =0`$. One has: $`\mathrm{sin}\delta _k=2k/m_{dec}1`$. (For example: If $`m_{dec}=3`$, then $`\delta _{1,2}=\pm 19.47`$ degree; if $`m_{dec}=4`$, then $`\delta _{1,3}=\pm 30.00`$ degree and $`\delta _2=0.00`$ degree; etc.) We also separate the sky in right ascension $`\alpha `$ into $`m_{ra}>1`$ stripes. They are defined by boundaries $`\alpha =360k^{}/m_{ra}`$ degree; $`k^{}=0,1,..,m_{ra}`$. (Obviously, a trivial modification of this separation in right ascension is the case, when the boundaries are $`\alpha =360(k^{}+p)/m_{ra}`$ degree, where $`p`$ is an arbitrary real number fulfilling $`0<p<1`$.) All this means that we separated the sky into $`M=m_{dec}\times m_{ra}`$ areas (”cells”) having the same size $`4\pi /M`$ steradian. If there are $`N`$ GRBs on the sky, then $`n=N/M`$ is the mean of GRBs at a cell. Let $`n_i;i=1,2,\mathrm{},M`$ be the observed number of GRBs at the $`i`$th cell ($`_{i=1}^Mn_i=N`$). Then $$var_M=(M1)^1\underset{i=1}{\overset{M}{}}(n_in)^2$$ (1) defines the observed variance. For the given cell structure with $`M`$ cells, due to the Bernoulli distribution (\[Mészáros 1997\], \[Balázs et al. 1998\]), the measured variance $`var_M`$ should be identical to the theoretically expected value $`n(11/M)`$. This theoretical prediction should then be tested. Note that this and similar methods (see, e.g., Mészáros (1997) for details and further references) are usual in astronomy. For example, this method was used already by Abell (1958) to reject the randomness in the sky-distribution of clusters of galaxies. The test is, compared with other statistical tests (”two-point angular correlation function”, ”nearest neighbor distances”, etc.; \[Peebles 1980\]; \[Diggle 1983\]; \[Pásztor 1993\]), not the most sensitive one to detect non-randomnesses. Its importance for our purposes is given by the fact that it allows an extremely simple generalization to the case with non-zero sky-exposure function. Now, we generalize the method to this case. This may easily be done by changing the boundaries of cells in order to have the same probability (and hence the same expected number $`n=N/M`$) for a given cell. The sky-exposure function is a function of declination only (\[Fishman et al. 1994\]; \[Meegan et al. 1998\]). Hence, the choice of equatorial coordinates is highly convenient, because then no changes of boundaries are necessary in right ascension. The new boundaries $`\delta _k,k=0,1,\mathrm{},m_{dec}`$ in declination may be calculated analytically as follows. Clearly, $`\delta _{0,m_{dec}}=\pm 90`$ degree remain. In BATSE Catalog \[Meegan et al. 1998\] the exposure function $`f(\delta )`$ is defined for 37 values of declination (for $`\delta _r=90,85,\mathrm{},+85,+90`$ degree; $`r=0,1,2,..,36`$). To obtain $`\delta _k`$ we, first, calculate the value $$A=\frac{5\pi }{180}\underset{r=1}{\overset{35}{}}f(\delta _r)\mathrm{cos}\delta _r,$$ (2) where for the given $`r`$ the corresponding declination is $`\delta _r=90+5r`$ degree. (Remark that $`r=0`$ and $`r=36`$, respectively, need not be in the sum, because $`\mathrm{cos}(\pm 90)=0`$.) Then, second, for $`m_{dec}2`$ we search for the values $`\delta _k;k=1,2,\mathrm{},(m_{dec}1)`$ as follows. For the given $`k`$ we search for the declination $`\delta _i`$ fulfilling the condition $$\frac{\pi }{36A}\underset{j=1}{\overset{i}{}}f(\delta _j)\mathrm{cos}\delta _j\frac{k}{m_{dec}}<\frac{\pi }{36A}\underset{j=1}{\overset{i+1}{}}f(\delta _j)\mathrm{cos}\delta _j.$$ (3) Having this we search, by linear interpolation between $`\delta _i`$ and $`\delta _{i+1}`$, for the exact value of $`\delta _k`$. By this method $`\delta _k`$ is well calculable. (For example: For $`m_{dec}=3`$ we obtain $`\delta _1=19.51`$ degree, $`\delta _2=22.44`$ degree; for $`m_{dec}=4`$ we obtain $`\delta _1=30.83`$ degree, $`\delta _2=1.51`$ degree, $`\delta _3=33.60`$ degree; etc.) Having these cells with these ”shifted” boundaries in declination the variance may be calculated identically to the case with no sky-exposure function. This method will test the pure intrinsic randomness; the effect of BATSE sky-exposure function is exactly eliminated. It is natural to probe different values of $`M`$. In addition, for some $`M`$ different cell structures are still possible (cf. $`M=12`$ allows $`m_{dec}=2,3,4,6`$). Hence, generally, several - say $`Q`$ \- cell structures may be probed for the same sample of GRBs. Having these $`Q`$ cell structures (and hence $`Q`$ means + $`Q`$ measured variances) two questions arise. 1. How to calculate the confidence level for a given cell structure? 2. Having $`Q`$ values of confidence levels, how to calculate the final confidence level? The answer for the first question seems to be quite clear: $`var_M/n`$ seems to be identical to the $`\chi ^2`$ value for $`M1`$ degree of freedom (\[Trumpler & Weaver 1953\]; \[Kendall & Stuart 1969\]; \[Press et al. 1992\]; the mean is obtained from the sample itself, and therefore the degree of freedom is $`M1`$). Nevertheless, the situation is not so obvious, because the $`\chi ^2`$ test needs $`n>5`$ (\[Trumpler & Weaver 1953\]; \[Kendall & Stuart 1969\]; \[Press et al. 1992\]). In addition, some statistical text-books propose to use ”quadratic” cells only (\[Diggle 1983\], Chapt. 2.5.). If all these restrictions were taken into account, then $`\chi ^2`$ tests would be possible only for $`2m_{dec}=m_{ra};M=2m_{dec}^2`$ and $`N>5M=10m_{dec}^2`$. This would be a drastic truncation of the possible cell structures. But, not doing these restrictions, the estimation of the confidence level for a given cell structure must be done by more complicated procedures; e.g., by numerical simulations. Concerning the answer to the second question the situation is even less clear. As the reasonable search for the final confidence level only Monte Carlo simulations seem to be usable \[Press et al. 1992\]. Keeping all this in mind, we will proceed as follows. In the coordinate system with axes $`x=1/M`$ versus $`y=\sqrt{var_M/n}=(var/mean)^{1/2}`$ the $`Q`$ values of $`(var/mean)^{1/2}`$ define $`Q`$ points (one point for any cell structure; $`y_j=\sqrt{var_{M,j}/n}`$, where $`j=1,2,\mathrm{}.,Q`$). Clearly, for these points one expects the theoretical curve $`y=\sqrt{1x}`$. This theoretical expectation can straightforwardly be verified, e.g., by least squares estimation (\[Press et al. 1992\], Chapt. 15.2.; \[Diggle 1983\], Chapt. 5.3.1). Our estimator is the dispersion $$\sigma _Q=\underset{j=1}{\overset{Q}{}}(y_j\sqrt{11/M})^2.$$ (4) Obviously, smaller $`\sigma _Q`$ suggests that the theoretical curve is better fitted. Note still that, as the best choice, the square root of $`var_M/n`$ is proposed in this ”var/mean” test (\[Diggle 1983\], Chapt. 5.4.). The confidence level can then be estimated by Monte Carlo simulations in the following way: We throw 1000-times randomly $`N`$ points on the sphere, and repeat the above calculation leading to $`\sigma _Q`$ for every simulated sample. Then we compare the size of the $`\sigma _Q`$ obtained from this simulation with $`\sigma _Q`$ obtained from the actual GRB positions. Let $`\omega `$ be the number of simulations, when the obtained $`\sigma _Q`$ is bigger than the actual value of $`\sigma _Q`$. Then one may conclude that $`(100\omega /10)`$ is the confidence level in percentage. Clearly, this method does not need $`n>5`$ and quadratic cells. There is no commonly accepted confidence level in statistics, above which the null-hypothesis should already be rejected (\[Trumpler & Weaver 1953\]; \[Kendall & Stuart 1969\]). It is only a general agreement that confidence levels smaller than $`95\%`$ should not be considered. Our opinion is (see also \[Kendall & Stuart 1969\]) that the confidence levels bigger than $`95\%`$ can already be taken as ”remarkable”, ”suspicious”, ”interesting”, etc.; a higher than 99% confidence level may still mean the rejection of null-hypothesis, and such result must doubtlessly be announced. Hence, we will require that the confidence level be bigger than $`95\%`$. Thus, here it must be $`\omega <50`$. In this paper GRBs will be taken between trigger-numbers 0105 and 6963 from Current BATSE Catalog \[Meegan et al. 1998\] having defined $`T_{90}`$ (i.e. all GRBs detected up to August 1996 having measured $`T_{90}`$). From them we exclude, similarly to Pendleton et al. (1997) and Balázs et al. (1998), the faintest GRBs having a peak flux (on 256 ms trigger) smaller than 0.65 photon/(cm<sup>2</sup>s). This truncation is proposed by Pendleton et al. (1997) in order to avoid the problems with the changing threshold. The 1284 GRBs obtained in this way define the ”all” class. From them there were 339 GRBs with $`T_{90}<2`$ s (the ”short” subclass), 181 GRBs with $`2`$ s $`<T_{90}<10`$ s (the ”intermediate” subclass) and 764 GRBs with $`T_{90}>10`$ s (the ”long” subclass). We will study the all class and the three subclasses separately. We will ad hoc choose $`m_{dec}=2,3,\mathrm{},8`$ and $`m_{ra}=2,3,\mathrm{},16`$. I.e. it will be $`Q=105`$. Of course, this choice of $`Q`$ is more or less subjective. Nevertheless, our choice is motivated by two concrete arguments. First, we would like to study only the angular scales much bigger than the positional errors. (The size of a cell will not be smaller than 22.5 degree. On these angular scales no problems should arise from the positional errors \[Meegan et al. 1998\].) Second, it is reasonable not to consider such high values of $`m_{dec}`$, when $`180/m_{dec}`$ is already comparable or even smaller than 5 degree. (If this were not required then the elimination of sky-exposure function would be problematic due to its definition for declination intervals with widths 5 degree.) ## 3 THE RESULTS Figure 1 collects the results of $`Q=105`$ ”var/mean” tests of four different cases. It is obvious immediately that for the ”all” case the points follow well the theoretical curve. For the ”short” and ”long” subclasses, on the other hand, there is a slight tendency of points to be above the theoretical curve. The situation concerning the intermediate subclass seems to be the most unambiguous: mainly for small $`M`$ (roughly below $`M40`$) the points are clearly above the theoretical curve. This suggests an intrinsic non-randomness mainly in the sky distribution of intermediate subclass; such possibility for the short and long subclasses, respectively, cannot be excluded, too. The results of Monte Carlo simulations support this expectation only in the case of intermediate subclass. We obtain $`\omega =287`$ ($`\omega =80`$, $`\omega =36`$, $`\omega =440`$) for all GRBs (short, intermediate, long GRBs). Hence, the rejection of null-hypothesis is confirmed for the intermediate subclass only on the 96.4% confidence level. For the short and long subclasses, respectively, and also for all GRBs the null-hypothesis cannot be rejected on the $`>95\%`$ confidence level. For the short subclass we have a $`92\%`$ confidence level; for the remaining two cases even smaller levels. ## 4 DISCUSSION AND CONCLUSION The most surprising result of paper concerns the intermediate subclass. The intrinsic non-randomness is confirmed on the confidence level $`>95\%`$. This confidence level, as discussed in Section 2, is ”remarkable”, but is not enough to reject the null-hypothesis of randomness. The results concerning the $`339`$ short GRBs should also be mentioned. Nevertheless, the $`92\%`$ confidence level is clearly not enough to reject the confidence null-hypothesis. On the other hand, this result, together with Balázs et al. (1998, 1999), suggest that also for the short subclass itself the rejection of null-hypothesis of intrinsic randomness can also occur by further tests. In the case of $`764`$ long GRBs, and also of the $`1284`$ all GRBs, there are no indications for the non-randomnesses. All this seems to be in accordance with the results of Balázs et al. (1998, 1999). We think that the result concerning the intermediate subclass is highly surprising, because just this new subclass, having the smallest number of GRBs, has a remarkable ”proper” behavior. A short further investigation of this subclass fully supports this conclusion. There are 181 GRBs in this intermediate subclass. Be divided this subclass into two further subclasses; into the ”dim” and ”bright” ones. By chance the peak flux = 2 photons/(cm<sup>2</sup>s) (on $`0.256`$s trigger) is practically identical to the medium of peak flux for this subclass. Therefore, we consider the GRBs having smaller (bigger) peak flux 2 photons/(cm<sup>2</sup>s) as the ”dim” (”bright”) subclass of the intermediate subclass. There are 92 GRBs at the ”dim” subclass, and 89 GRBs at the ”bright” one. We provide the 105 ”var/mean” tests for these two parts, too. We obtain the surprising result that the ”dim” subclass has an intrinsic non-randomness on the 99.3% confidence level ($`\omega =7`$). Contrary this, the ”bright” subclass can still be random ($`\omega =662`$). The sky distribution of $`92`$ intermediate dim GRBs is shown on Figure 2. We mean that the behavior of the intermerdiate subclass of GRBs, quite independently, supports the correctness of the introduction of this new subclass (\[Mukherjee et al. 1998\]; \[Horváth 1998\]). Further investigations of this new subclass are highly required. Three notes are still needed. First, purely from the statistical point of view, it must be precised that even the rejection of null-hypothesis of the intrinsic randomness would not mean a pure intrinsic non-randomness in the spatial angular distribution of GRBs. This is given by the fact that, up to now, it cannot be fully excluded that GRBs (or some part of them) are not unique phenomenons, and there can occur some repetitions, too. This question is studied intensively by several papers (\[Meegan et al. 1995\], \[Quashnock 1995\], \[Quashnock 1996\], \[Tegmark et al. 1996a\], \[Graziani et al. 1998\], \[Hakkila et al. 1998\]) concluding that repetition can still play a role. Second, strictly speaking, the statistical counts-in-cells test is testing the ”complete spatial randomness” (shortly the ”randomness”) of the distribution of GRB on the celestial sphere (\[Diggle 1983\], Chapt.1.3). Therefore, in this paper we have kept this terminology. In cosmology, on the other hand, the word ”random” (”non-random”) is rarely used, and the word ”isotropic” (”anisotropic”) is usual (for the exact definition of isotropy in cosmology see, e.g., Weinberg (1972), Chapt. 14.1). Of course, here we will not go into the details of these terminology questions (see, e.g., Peebles (1980) for more details concerning these questions). We note only that the ”random-isotropic” (”non-random-anisotropic”) substitution is quite acceptable on the biggest angular scales; on smaller angular scales the situation is not so clear. Therefore, in Balázs et al. (1998, 1999), where only the angular scales $`90`$ degrees and higher were studied, the words ”isotropy” and ”anisotropy” were quite usable. In this article, going down up to the scales $`(2025)`$ degree, the used terminology is more relevant. Third, trivially, further studies are needed. They should test - by other different statistical methods - again the intrinsic randomnesses (more generally: the intrinsic spatial distributions \[Lamb 1997\]), both for all GRBs and for the subclasses. In addition, a test of the repetition alone, i.e. a test not being influenced by positions, is highly required. As the conclusion, the results of paper may be summarized as follows. * We developed a method, which can verify quite simply the intrinsic randomness alone in the angular distribution of GRBs, because the method eliminates exactly the non-zero sky-exposure function. * We rejected the null-hypothesis of the intrinsic randomness in the angular distribution of 181 intermediate GRBs on the $`96.4\%`$ confidence level. * We rejected the null-hypothesis of the intrinsic randomness in the angular distribution of 92 ”dim” intermediate GRBs on the $`99.3\%`$ confidence level. * We did not reject the null-hypotheses of the intrinsic randomnesses in the angular distribution of the remaining two subclasses and of the all GRBs, respectively, on the $`>95\%`$ confidence levels; the ”bright” intermediate GRBs seem to be distributed randomly, too. We thank the valuable discussions with Drs. Michael Briggs, Peter Mészáros, László Pásztor, Dennis Sciama, Gábor Tusnády and anonymous referee. One of us (A.M.) thanks for the hospitality at Konkoly Observatory and Eötvös University. This article was partly supported by GAUK grant 36/97, by GAČR grant 202/98/0522, by Domus Hungarica Scientiarium et Artium grant (A.M.), by OTKA grant T024027 (L.G.B) and by OTKA grant F029461 (I.H.). ## FIGURES Figure 1. The results of $`105`$ “var/mean” tests of four different cases drawn in the $`1/M`$ vs. $`\sqrt{\text{var/mean}}`$ frame. The theoretical curve $`\sqrt{11/M}`$ (solid line) is also shown. $`M`$ is the number of cells. Figure 2. Sky distribution of 92 GRBs of ”dim” subclass of the intermediate subclass in equatorial coordinates.
warning/0003/cond-mat0003308.html
ar5iv
text
# Nonequilibrium thermodynamics of colloids ## I Introduction Much is known about the hydrodynamic description of colloidal systems. This is described in a number of well written books on the subject . In these books a multitude of details about the interesting behavior of suspended asymmetric colloidal particles in shear flow is treated. What is not being done is to treat these systems in the context of nonequilibrium thermodynamics. That such a treatment is possible becomes clear reading the monograph by De Gennes on The Physics of Liquid Crystals. Rather than going into detail regarding the motion of individual particles he discusses, both the equilibrium and the nonequilibrium properties of these materials, using thermodynamics. In this paper we will apply nonequilibrium thermodynamics to these colloidal systems. As it is not our intention to duplicate all the well described knowledge about the collective behavior of these systems, we will focus on one aspect which has remained puzzling from the viewpoint of nonequilibrium thermodynamics. It is well-known that the viscosity in these systems depends on the shear rate. In view of the fact that the shear rate is not a thermodynamic state variable, like, for instance, the temperature, classical nonequilibrium thermodynamics, as explained in the classic monograph by de Groot and Mazur , does not allow such a dependence. Books on colloidal hydrodynamics explain this dependence in terms of the behavior of asymmetric suspended particles in shear flow. The question we pose ourselves in this paper is: explain this dependence using nonequilibrium thermodynamics alone. Interestingly enough Prigogine and Mazur already mentioned this problem as one of the three possible examples of their method of internal variables. The orientation of the suspended particles was suggested as such an internal variable. In de Groot and Mazur’s book the other two examples are treated in detail but this one was not pursued. It had to wait for a better understanding of these systems. In this paper we will show that it is now possible to answer this question for a suspension of asymmetric colloidal particles. In addition to answering a question of principle, the treatment is simpler and more direct than the current explanations. It is also not restricted to dilute suspensions. In the second section we give the Gibbs relation for an isotropic system with asymmetric colloidal particles. The particles are assumed to have a symmetry axis. A new term appears which gives the entropy change due to a change in a tensorial order parameter $`𝐐`$. This order parameter characterizes the alignment of the particles . This is precisely the new internal variable Prigogine and Mazur needed. In the third section the resulting entropy production is presented. This leads to linear laws in which the viscous pressure tensor and the time rate of change of the order parameter are expressed in their conjugate thermodynamic forces. In the fourth section these laws are solved for the special case of stationary shear flow. An expression is found for the viscous pressure in terms of the shear gradient. Their ratio defines an effective viscosity. This effective viscosity is found to be a function of the shear. Crucial to obtain this result is the fact that the Onsager coefficients in general depend on the order parameter $`𝐐`$ which is a state variable. This answers the principle question. We proceed in chapter five with a analysis of time dependent shear flow. It is found that this results in a frequency dependent effective shear viscosity. In the last section we derive a simple expression for the dependence of one of the Onsager coefficients on the order parameter to make this point clear beyond doubt. ## II The Gibbs relation In an isotropic suspension of particles, which are cylindrically symmetric, a tensorial order parameter $`𝐐`$ appears as a new thermodynamic variables . $`𝐐`$ is a traceless symmetric tensor, which is defined as the ensemble average of $`\left(\mathrm{𝐚𝐚}\mathrm{𝟏}/3\right),`$ where $`𝐚`$ is a unit vector along the symmetry axis of one of the particles and $`\mathrm{𝟏}`$ is the unit tensor. In an isotropic phase $`𝐐`$ is, in the absence of any directing field, zero. In, for instance, a magnetic field $`𝐐`$ is nonzero. Changes of $`𝐐`$ lead to changes of the entropy. It is therefore important to take the order parameter along in the description of the system. The total differential of the entropy is given by the following Gibbs relation $$Tds_v=du_v\mu d\rho \mu _sd\rho _s𝐰:d𝐐$$ (1) Here $`T`$ is the temperature, $`s_v=\rho s`$ the entropy density, $`u_v=\rho u`$ the internal energy density, $`\mu `$ the chemical potential of the solvent, $`\rho `$ the solvent particle number density, $`\mu _s`$ the chemical potential of the suspended particles and $`\rho _s`$ the suspended particle number density. The densities with a subscript $`v`$, in addition to $`\rho `$ and $`\rho _s,`$ are per unit of volume. $`𝐰`$ is the variable conjugate to $`𝐐`$ and is given by $$𝐰=T\left(\frac{s_v}{𝐐}\right)_{u_v,\rho ,\rho _s}$$ (2) Alternatively one may use the relation to the free energy $`f_v`$ $$𝐰=\left(\frac{f_v}{𝐐}\right)_{T,\rho ,\rho _s}$$ (3) Furthermore $`:`$ between two matrices signifies a double contraction. De Gennes \[4 (page 48),5\] writes as explicit expression for the free energy density $$f_v(T,\rho ,\rho _s,𝐐)=f_v^0(T,\rho ,\rho _s)+\frac{1}{2}A(T,\rho ,\rho _s)tr(𝐐.𝐐)+\frac{1}{3}B(T,\rho ,\rho _s)tr(𝐐.𝐐.𝐐)+\left(𝐐^4\right)$$ (4) where $``$ is the order symbol. Here $`tr`$ signifies the trace of a matrix. Note that $`tr(𝐐.𝐐)=𝐐:𝐐`$. The resulting traceless symmetric conjugate variable is $$𝐰=A(T,\rho ,\rho _s)𝐐+B(T,\rho ,\rho _s)\stackrel{}{\overline{𝐐.𝐐}}+\left(𝐐^3\right)$$ (5) where $`\stackrel{}{\overline{𝐐.𝐐}}`$ is the traceless symmetric part of $`𝐐.𝐐`$. This relation may be inverted to give the order parameter in terms of its conjugate variable $$𝐐=A^1(T,\rho ,\rho _s)𝐰A^3(T,\rho ,\rho _s)B(T,\rho ,\rho _s)\stackrel{}{\overline{𝐰.𝐰}}+\left(𝐰^3\right)$$ (6) The typical size of $`A`$ is of the order of $`k_BT\rho _s`$ where $`k_B`$ is Boltzmann’s constant. ## III The entropy production Without the $`𝐐`$ term the Gibbs relation is equivalent to the one used in de Groot and Mazur . Note that they use densities per unit of mass while we use densities per unit of volume like de Gennes is doing. The calculation of the entropy production, using the balance equations, is essentially identical to the one they give. The only difference is the appearance of a term due to the order parameter $`𝐐`$. In order to describe the contributions of the order parameter to the dynamical behavior of the system we may restrict ourselves, using the Curie principle, to contributions of force-flux pairs which are symmetric traceless tensors. The rate of entropy production density, as a function of position $`𝐫`$ and time $`t`$, due to these contributions becomes $$T\sigma _{tens}(𝐫,t)=\stackrel{}{\overline{𝚷(𝐫,t)}}:\stackrel{}{\overline{grad𝐯(𝐫,t)}}𝐰(𝐫,𝐭):\frac{d𝐐(𝐫,𝐭)}{dt}$$ (7) here $`\stackrel{}{\overline{𝚷}}`$ is the symmetric traceless part of the viscous pressure tensor and $`\stackrel{}{\overline{grad𝐯}}`$ the symmetric traceless part of the velocity gradient. The resulting linear laws are $`\stackrel{}{\overline{𝚷(𝐫,t)}}`$ $`=`$ $`2\eta \stackrel{}{\overline{grad𝐯(𝐫,t)}}\mathrm{}𝐰(𝐫,𝐭)`$ (8) $`{\displaystyle \frac{d𝐐(𝐫,𝐭)}{dt}}`$ $`=`$ $`\mathrm{}\stackrel{}{\overline{grad𝐯(𝐫,t)}}\mathrm{}_Q𝐰(𝐫,𝐭)`$ (9) Here we used that, $`\stackrel{}{\overline{grad𝐯(𝐫,t)}}`$ is odd and $`𝐰(𝐫,𝐭)`$ even for time reversal,so that the matrix of Onsager coefficients is antisymmetric. The Onsager coefficients are in general functions of the thermodynamic state variables, so that we have $$\eta (T,\rho ,\rho _s,𝐐),\mathrm{}(T,\rho ,\rho _s,𝐐)\text{and}\mathrm{}_Q(T,\rho ,\rho _s,𝐐)$$ (10) In line with the assumption of linearity, they are not allowed to be functions of $`\stackrel{}{\overline{grad𝐯(𝐫,t)}}`$ and $`d𝐐(𝐫,𝐭)/dt`$. The diagonal coefficient $`\eta `$ is the viscosity and the diagonal coefficient $`\mathrm{}_Q`$ is a typical rotational mobility of the suspended particles divided by $`\rho _s`$. The cross coefficient $`\mathrm{}`$ is dimensionless and is expected to be a positive constant of the order unity. ## IV Stationary elongational flow The above equations may easily be solved for stationary elongational flow $$𝐯(𝐫,t)=\frac{1}{2}\stackrel{.}{\gamma }(y,x,0)$$ (11) The resulting velocity gradient is $$\stackrel{}{\overline{grad𝐯(𝐫,t)}}=grad𝐯(𝐫,t)=\frac{1}{2}\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)$$ (12) Using the second linear law, eq.(9b), results in $$𝐰(𝐫,𝐭)=\frac{\mathrm{}}{\mathrm{}_Q}\stackrel{}{\overline{grad𝐯(𝐫,t)}}=\frac{\mathrm{}}{2\mathrm{}_Q}\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)$$ (13) Substitution in eq.(6) gives $$𝐐=\frac{\mathrm{}}{2A\mathrm{}_Q}\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)+\left(\stackrel{2}{\stackrel{.}{\gamma }}\right)$$ (14) The first linear law, eq.(9a), then yields $$\stackrel{}{\overline{𝚷(𝐫,t)}}=(\eta +\frac{\mathrm{}^2}{2\mathrm{}_Q})\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)\eta _{eff}\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)$$ (15) Writing the dependence on the thermodynamic state variables explicitly, the effective viscosity becomes $$\eta _{eff}(T,\rho ,\rho _s,𝐐)=\eta (T,\rho ,\rho _s,𝐐)+\frac{\mathrm{}^2(T,\rho ,\rho _s,𝐐)}{2\mathrm{}_Q(T,\rho ,\rho _s,𝐐)}$$ (16) In view of the above relation of the order parameter with the rate of elongation, one may write this relation as $$\eta _{eff}(T,\rho ,\rho _s,\stackrel{.}{\gamma })=\eta (T,\rho ,\rho _s,\stackrel{.}{\gamma })+\frac{\mathrm{}^2(T,\rho ,\rho _s,\stackrel{.}{\gamma })}{2\mathrm{}_Q(T,\rho ,\rho _s,\stackrel{.}{\gamma })}$$ (17) The effective viscosity depends therefore on the rate of elongation. ## V Oscillatory elongational flow The above equations may also easily be solved in the case of oscillatory elongational flow. In that case the velocity field, its symmetric traceless gradient, the order parameter and its conjugate variable all become proportional to $`\mathrm{exp}(i\omega t)`$. We now neglect the nonlinear terms in eq.(5) and the $`𝐐`$ dependence of the transport coefficients in order to avoid the appearance of higher order harmonics. Though this is an interesting phenomenon, it is not our present concern. In the second linear law we must now replace $`d𝐐(𝐫,𝐭)/dt`$ by $`i\omega 𝐐(𝐫,𝐭)`$. In the solution this implies that $$𝐐=A^1𝐰=\frac{\mathrm{}}{A\mathrm{}_Qi\omega }\stackrel{}{\overline{grad𝐯}}$$ (18) The resulting complex frequency dependent effective viscosity then becomes $$\eta _{eff}(T,\rho ,\rho _s,\omega )=\eta (T,\rho ,\rho _s)+\frac{\mathrm{}^2(T,\rho ,\rho _s)}{2\mathrm{}_Q(T,\rho ,\rho _s)\left(1i\omega \tau (T,\rho ,\rho _s)\right)}$$ (19) with a relaxation time $$\tau (T,\rho ,\rho _s)=\frac{1}{A(T,\rho ,\rho _s)\mathrm{}_Q(T,\rho ,\rho _s)}$$ (20) This relaxation time is the reorientation time of the suspended particles due to rotational diffusion, and is therefore of the order of the rotational diffusion coefficient. It follows that $`\omega \tau `$ is the relevant Péclet number for rotational diffusion in an oscillating velocity field. For low frequencies the effective viscosity is enhanced by the coupling to the order parameter $$\eta _{eff}\left(T,\rho ,\rho _s,\omega =0\right)=\eta (T,\rho ,\rho _s)+\frac{\mathrm{}^2(T,\rho ,\rho _s)}{2\mathrm{}_Q(T,\rho ,\rho _s)}$$ (21) This is the same result as found above, eq.(16), for the stationary case if one neglects the dependence of the coefficient on the order parameter. For high frequencies one finds $$\eta _{eff}\left(T,\rho ,\rho _s,\omega =\mathrm{}\right)=\eta (T,\rho ,\rho _s)$$ (22) Writing the complex viscosity as the sum of a real and imaginary part, $`\eta _{eff}=\eta _{eff}^{}+i\eta _{eff}^{\prime \prime }`$, one has $$\eta _{eff}^{}(T,\rho ,\rho _s,\omega )=\eta (T,\rho ,\rho _s)+\frac{\mathrm{}^2(T,\rho ,\rho _s)}{2\mathrm{}_Q(T,\rho ,\rho _s)\left(1+\omega ^2\tau ^2(T,\rho ,\rho _s)\right)}$$ (23) and $$\eta _{eff}^{\prime \prime }(T,\rho ,\rho _s,\omega )=\frac{\omega \mathrm{}^2(T,\rho ,\rho _s)}{2\mathrm{}_Q(T,\rho ,\rho _s)\left(1+\omega ^2\tau ^2(T,\rho ,\rho _s)\right)}$$ (24) This last expression gives the elastic contribution to the complex viscosity. ## VI The Onsager coefficients Nonequilibrium thermodynamics does not derive expressions for the Onsager coefficients. Like the thermodynamic derivatives of the entropy as for instance $`A(T,\rho ,\rho _s)`$, they have to be either measured or calculated from a description in terms of the motion of the separate particles. Above we have given estimates of the typical order of magnitude of the various coefficients on the basis of the physical meaning of the coefficients and dimensional arguments. As the shear dependence of the effective viscosity is the property we set out to find, we will pursue this element a bit further. One may write $$\eta =\eta _0\left(1+\left[\eta \right]\varphi \right)$$ (25) where $`\varphi =V\rho _s,`$ with $`V`$ the volume of the monodisperse suspended particles. $`\left[\eta \right]`$ is the so-called intrinsic viscosity. Each suspended particle contributes to the viscous dissipation. For low densities the intrinsic viscosity is independent of their concentration. Einstein found that the intrinsic viscosity for spheres was equal to 2.5. For rods the contribution depends on the orientation $`𝐚`$ of the rod. In order to get the intrinsic viscosity one must then first calculate the contribution of a rod to the viscous pressure tensor as a function of its orientation and then average over the orientations . The distribution over orientations may be calculated using the rotational diffusion equation for this distribution. For the stationary state distribution corresponding to a given order parameter $`𝐐`$ one has $$f\left(𝐚\right)=\frac{15}{8\pi }\left(\mathrm{𝐚𝐚}\frac{\mathrm{𝟏}}{3}\right):𝐐+\frac{1}{4\pi }$$ (26) For spheroids, which are useful as a model system, the solution of the velocity field has been given by Jeffery . In the monograph by van de Ven the resulting contribution to the pressure tensor is given. In the appendix we average those expressions over the above orientation distribution. This results in $`\left[\eta \right]`$ $`=`$ $`{\displaystyle \frac{2}{15}}(r_e^21)[{\displaystyle \frac{5\mathrm{}}{4\left(2r_e^2\left(12r_e^2\right)A_e\right)}}{\displaystyle \frac{\stackrel{.}{\gamma }}{A\mathrm{}_Q}}`$ (28) $`+{\displaystyle \frac{26r_e^224r_e^2A_e15A_e}{\left(2r_e^23A_e\right)\left(2r_e^2\left(2r_e^2+1\right)A_e\right)}}+{\displaystyle \frac{6}{\left(r_e^2+1\right)\left(3A_e2\right)}}]`$ Here the aspect ratio $`r_e=a/b`$ is the ratio of the diameter $`a`$ along the symmetry axis and the diameter $`b`$ normal to the symmetry axis. Furthermore $`A_e`$ $`=`$ $`{\displaystyle \frac{r_e^2}{r_e^21}}{\displaystyle \frac{r_earccoshr_e}{\left(r_e^21\right)^{3/2}}}\text{for }r_e>1`$ (29) $`A_e`$ $`=`$ $`{\displaystyle \frac{r_e\mathrm{arccos}r_e}{\left(1r_e^2\right)^{3/2}}}{\displaystyle \frac{r_e^2}{1r_e^2}}\text{for }r_e<1`$ (30) Note that $`\stackrel{.}{\gamma }/A\mathrm{}_Q`$ is the rotational Péclet number so that eq.(28) is an expansion in the Péclet number. In the limit of a sphere, $`r_e1,`$ one may verify that the expression reduces to 2.5. In that case the shear dependence of the effective viscosity disappears. Prolate spheroids, $`r_e>1`$, are shear thickening and oblate spheroids, $`r_e<1`$, are shear thinning for small Péclet numbers in elongational flow . ## VII Discussion and conclusions The main objective was to show that a straightforward application of nonequilibrium thermodynamics leads to a shear dependent viscosity. We were able to derive a general nonlinear expression for this dependence. The important reason for this is the coupling to the orientation of the suspended colloidal particles. This orientation is described by a tensorial order parameter $`𝐐`$ as thermodynamic state variable. The fact that the Onsager coefficients depend on this state variable then results in a shear dependent viscosity. Of course the critical reader could have doubted such a dependance of the Onsager coefficients on $`𝐐.`$ In order to defend us against this critique, we derive an explicit formula for small Péclet numbers, $`\stackrel{.}{\gamma }/A\mathrm{}_Q`$. This gives a linear relation for this dependence. For larger Péclet numbers the nonlinear relation between $`𝐐`$ and $`𝐰`$ must be taken into account. This will lead to more realistic and interesting predictions. One interesting aspect is that the powers of $`𝐐`$ lead to shear field with a different ”direction”. As the principle question, why the viscosity depends on the shear, has been answered, we will not here pursue these points. The frequency dependence is due to a single exponential decay of the order parameter. For low frequencies the viscosity is enhanced by the coupling to the orientation of the colloidal particles. For higher frequencies the orientation lags behind. This leads to thinning of the fluid and a reversible elastic response. One can extend the analysis, within the general frame of nonequilibrium thermodynamics, by introducing the whole angular distribution of the particles as an internal thermodynamic variable . The details for such an analysis have recently been worked out by Mazur . This results in a whole spectrum of relaxation times and consequently gives a more complex frequency dependence in the transition from low frequency to high frequency behavior. In our analysis we do not need to analyse the orientational distribution function in any detail. The knowledge of the order parameter is enough. This makes our analysis much simpler that the usual procedure, where the rotational diffusion equation for this distribution function is solved numerically. As we said above , the orientational distribution function can be introduced as an internal thermodynamic variable. In addition to a richer frequency dependence this will also lead to a more complex shear dependence. The low frequency and the low shear behavior of the viscosity has a very similar non-analytic dependance on these parameters . This is a general phenomenon which is not specifically due to the asymmetry, or for that matter the presence, of the suspended particles. It falls outside the scope of the present article. In the field of socalled extended irreversible thermodynamics , one assumes that fluxes, fluxes of fluxes, etc., are also thermodynamic state variables. This assumption distinguishes it from classical irreversible thermodynamics, where this is not considered to be a proper assumption. The viscous pressure is such a flux and is, according to extended irreversible thermodynamics, a state variable. On the basis of this assumption the viscosity can be a function of the viscous pressure. As a result, the viscosity then becomes a function of the shear. The assumption, so to say, defines the problem away. We will not try to invalidate their assumption. We only want to stress that the present paper shows that such an assumption is not needed to obtain a shear dependent viscosity. As such their assumption seems to be one to many. Acknowledgement We want to thank T.G.M. van de Ven for a clarification of some of the formulae in his book. D. Bedeaux wants to thank DGICYT for supporting a visit to the University of Barcelona where this work was started. J.M. Rubi wants to thank N.W.O. for support of a visit to Leiden University where part of this work was finalized. ## VIII Appendix We use the results of Jeffery as presented in van de Ven’s monograph , see in particular pages 229-231. The shear field he uses is $`v_1=v_2=0`$ and $`v_3=Gx_2`$. In our notation $`x=x_2,y=x_3,z=x_1,\varphi =c`$ and $`\stackrel{.}{\gamma }=G`$. The symmetric traceless part of this simple shear field is identical to the one we used above. In view of the linearity of the problem we may therefore use the symmetric traceless part of the contribution to the viscous pressure tensor van de Ven uses: $$𝚷=𝚷_0+\frac{8\pi \eta _0}{V}\varphi 𝐌$$ (31) where $`V=4\pi b^3r_e/3`$ is the volume of the particle. $`𝚷_0=2\eta _0grad𝐯(𝐫,t)`$ is the viscous pressure one would have in the absense of the suspended particles. He gives the tensor $`𝐌`$ as $`𝐌^{}`$ in eq.(3.280) in a coordinate frame attached to the particle. Using the distribution given above, eq.(26), we have $$𝐌=\frac{15}{8\pi }\left[𝐌\left(𝐚\right)\left(\mathrm{𝐚𝐚}\frac{\mathrm{𝟏}}{3}\right)𝑑\mathrm{\Omega }\right]:𝐐+\frac{1}{4\pi }𝐌\left(𝐚\right)𝑑\mathrm{\Omega }$$ (32) where $`d\mathrm{\Omega }=\mathrm{sin}\theta d\theta d\varphi `$ with $`0<\theta <\pi `$ and $`0<\varphi <2\pi `$. The integral between square brackets gives a constant times the four index symmetric traceless unit tensor $`𝚫`$. The matrix elements of this tensor are given by $$\mathrm{\Delta }_{ijkl}=\frac{1}{2}\delta _{ik}\delta _{jl}+\frac{1}{2}\delta _{il}\delta _{jk}\frac{1}{3}\delta _{ij}\delta _{kl}$$ (33) The constant can be calculated by taking $`i=l`$, $`j=k`$ and summing over $`i`$ and $`j`$. This results in $$𝐌=\frac{3}{16\pi }[𝐌\left(𝐚\right):(\mathrm{𝐚𝐚}\frac{\mathrm{𝟏}}{3})d\mathrm{\Omega }]𝐐+\frac{1}{4\pi }𝐌\left(𝐚\right)d\mathrm{\Omega }$$ (34) The contraction of two matrices is independent of the reference frame and we may therefore use the reference frame attached to the particle. In this reference frame $`\left(\mathrm{𝐚𝐚}\mathrm{𝟏}/3\right)=\left(2/3\right)𝐞_1𝐞_1\left(1/3\right)𝐞_2𝐞_2\left(1/3\right)𝐞_3𝐞_3`$ in terms of the unit vectors in this frame. This gives $$𝐌=\frac{1}{16\pi }\left[\left(2M_{11}^{}M_{22}^{}(\theta ,\varphi )M_{33}^{}(\theta ,\varphi )\right)𝑑\mathrm{\Omega }\right]𝐐+\frac{1}{4\pi }𝐌(\theta ,\varphi )𝑑\mathrm{\Omega }$$ (35) Using eq.(3.280) we see that $`M_{22}^{}`$ and $`M_{33}^{}`$ integrate to zero while $`M_{11}^{}`$ is independent of direction. In the second integral we can use eqs.(3.282)-(3.284) to obtain the only nonzero contribution. The result is $$𝐌=\frac{1}{2}M_{11}^{}𝐐+\frac{b^3r_e}{6}\left(\frac{4}{15}C_1+\frac{1}{3}C_2+\frac{2}{3}C_3\right)\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)$$ (36) Substitution of $`𝐐`$ then yields $$𝐌=\left[\frac{\mathrm{}}{4A\mathrm{}_Q}M_{11}^{}+\frac{b^3r_e}{6}\left(\frac{4}{15}C_1+\frac{1}{3}C_2+\frac{2}{3}C_3\right)\right]\stackrel{.}{\gamma }\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)$$ (37) Substitution into eq.(31) and comparing with eq.(25) gives for the intrinsic viscosity $$\left[\eta \right]=\left[\frac{\mathrm{}}{A\mathrm{}_Q}\frac{3M_{11}^{}}{2b^3r_e}+\frac{4}{15}C_1+\frac{1}{3}C_2+\frac{2}{3}C_3\right]$$ (38) Substitution of eqs.(3.280) and (3.284) finally gives eq.(28).
warning/0003/hep-ph0003063.html
ar5iv
text
# Lepton flavor violation in the supersymmetric standard model with vectorlike leptons ## I Introduction The $`\mathrm{E}_6`$ supersymmetric (SUSY) grand unified theory (GUT) and the $`\mathrm{E}_8\times \mathrm{E}_8^{}`$ superstring theory are good candidates for unified theories of elementary particle physics. While such theories manifest themselves at the ultrahigh energy scale ($`10^{16}`$ GeV), it is natural to expect that they have some effects on physics at the electroweak scale ($`10^2`$ GeV) e6 . It is, however, difficult to determine precisely what effective theory actually appears at the electroweak scale. This may be understood by considering the facts that the $`\mathrm{E}_6`$ SUSY GUT has many possible symmetry breaking patterns and that the superstring theory has many possible compactification patterns. Hence, in the present experimental circumstances, it will not be so useful to perform detailed investigations on specific models. We are rather interested in the general aspects of the low-energy effects of the unified theories. Particularly, in this article we study the effects on lepton physics at the electroweak scale. In the standard model, the lepton number and flavor are conserved automatically. Hence, if the processes violating these quantum numbers are observed, including the neutrino masses and oscillations, $`\mu e\gamma `$, $`\mu 3e`$, and so on, they indeed provide important information of new physics above the electroweak scale. In the supersymmetric models with supergravity scenario Chamseddine:1982jx , which would be derived from the unified theories, the renormalization effects on the soft supersymmetry breaking terms may induce sizable lepton flavor violation. These lepton flavor violating renormalization effects are actually obtained in the SUSY GUT Barbieri:1994pv ; Barbieri:1995tw ; hisano2 and the minimal supersymmetric standard model (MSSM) with heavy right-handed neutrinos Borzumati:1986qx ; hisano ; Hisano:1998fj . Here, as another intriguing possibility for lepton flavor violation, we investigate the supersymmetric standard model incorporating vectorlike $`\mathrm{SU}(2)_\mathrm{L}`$ doublet leptons ll ; Dubovsky:1997rq . Especially, the significant slepton mixing may be generated by radiative corrections with the couplings between the ordinary leptons and these exotic leptons. It is remarkable that this new physics with exotic leptons can be discovered just around the electroweak scale in contrast to the cases of GUT and right-handed neutrinos. The present model is motivated by the $`\mathrm{E}_6`$ type unification. The $`\mathrm{E}_6`$ 27 representation contains one generation of the ordinary quarks and leptons in the standard model, the vectorlike $`\mathrm{SU}(2)_\mathrm{L}`$ singlet quarks, the vectorlike $`\mathrm{SU}(2)_\mathrm{L}`$ doublet leptons, the right-handed neutrino, and the $`\mathrm{SU}(2)_\mathrm{L}`$ singlet Higgs field. The ordinary left-handed neutrinos may acquire Majorana masses through the seesaw mechanism with the right-handed neutrinos seesaw . These ordinary neutrino masses can be very small as required phenomenologically, if the right-handed neutrinos are superheavy or if they have extremely small Yukawa couplings with the left-handed lepton doublets. Hence, given the small ordinary neutrino masses, we do not consider further the effects of the right-handed neutrinos in the electroweak physics. In this investigation, we concentrate on the lepton flavor violating effects of the vectorlike lepton doublets $`(L,L^c)`$ and their scalar superpartners. Specifically, these supersymmetric effects of the exotic leptons may provide the significant Br($`\mu e\gamma `$), while the lepton flavor changing neutral currents (FCNC’s) at the tree level are suppressed naturally by the hierarchy of the charged lepton masses. The down type quark singlets $`(D^c,D)\overline{\mathrm{𝟑}}_C+\mathrm{𝟑}_C`$ are also present in the $`\mathrm{E}_6`$ unification, which may form the $`\overline{\mathrm{𝟓}}+\mathrm{𝟓}`$ of SU(5) subgroup with $`(L,L^c)\mathrm{𝟐}_\mathrm{L}+\mathrm{𝟐}_\mathrm{L}`$. The survival of these exotic particles to the electroweak scale depends on the symmetry breakings. The $`\mathrm{𝟐}_\mathrm{L}`$-$`\mathrm{𝟑}_C`$ mass splitting in the 5 can be realized with suitable Higgs multiplets, e.g. the 351 of $`\mathrm{E}_6`$ including the 24 of SU(5), as usually considered to obtain the light $`\mathrm{SU}(2)_\mathrm{L}`$ doublet Higgs field. Hence, it is possible that $`(L,L^c)`$ survive while $`(D^c,D)`$ become superheavy in the 27. It may also be considered in the superstring theory that $`(L,L^c)`$ become light while $`(D^c,D)`$ do not in the $`\overline{\mathrm{𝟐𝟕}}+\mathrm{𝟐𝟕}`$. On the other hand, $`(D^c,D)`$ and $`(L,L^c)`$ are expected to survive together for the gauge coupling unification. Even in this case, $`(D^c,D)`$ and $`(L,L^c)`$ may originate in different 27’s and/or ($`\overline{\mathrm{𝟐𝟕}}+\mathrm{𝟐𝟕}`$)’s. Then, it should be noted in these cases, whether $`(D^c,D)`$ survive or not, that the simple $`\mathrm{E}_6`$ relations no longer hold for the Yukawa couplings after the $`\mathrm{E}_6`$ breaking. In particular, the flavor violating couplings with $`(L,L^c)`$ may be described independently of those with $`(D^c,D)`$. Given this general situation, we concentrate on the lepton flavor violation with $`(L,L^c)`$ in the present article. A detailed examination of the quark flavor violation with $`(D^c,D)`$, particularly the supersymmetric effects, is reserved for a separate publication. (Such supersymmetric effects are investigated in Ref.Dubovsky:1997rq in a context different from the present one.) Of course, if $`(D^c,D)`$ and $`(L,L^c)`$ originate in the same 27, the flavor violating effects of these exotic leptons and quarks will be provided from some common $`\mathrm{E}_6`$ coupling. In this specific case, the calculations made in the present analysis for the lepton flavor violation such as $`\mu e\gamma `$ can be extended readily for the quark flavor violation such as $`K^0`$-$`\overline{K}^0`$ mixing. Simple order estimates, as will be quoted occasionally in the text, indicate that the constraints from the quark flavor violating effects are in fact less stringent than those from the lepton flavor violating effects. This paper is organized as follows. In Section II, we describe the lepton mixing in the presence of vectorlike lepton doublets. Then, the contributions to Br($`\mu e\gamma `$) due to the ordinary-exotic lepton mixing are shown to be negligibly small. In Section III, we investigate the slepton mixing generated by radiative corrections with the ordinary-exotic lepton couplings. Significant contributions to Br($`\mu e\gamma `$) can be obtained from this slepton mixing. Section IV is devoted to the summary. ## II The lepton mixing In this section, we describe the lepton mixing including the vectorlike lepton doublets, which contributes to the lepton flavor changing processes such as $`\mu e\gamma `$. We show that the modification of the gauge interactions provided by this lepton mixing is fairly suppressed even if the relevant flavor violating Yukawa couplings take significant values. This feature is indeed attributed to the mass hierarchy between ordinary leptons and exotic leptons. The ordinary leptons are represented by $`l_i(\nu _i,e_i)`$ and $`e_i^c`$ with the subscript $`i=1,2,3`$ for the generations. The vectorlike leptons and singlet Higgs field are also listed as follows with their quantum numbers of the $`\mathrm{SU}(2)_\mathrm{L}\times \mathrm{U}(1)_Y`$: $`Ll_4=\left(\begin{array}{c}N\\ E\end{array}\right):(\mathrm{𝟐},1/2),L^c=\left(\begin{array}{c}E^c\\ N^c\end{array}\right):(\mathrm{𝟐},1/2),S:(\mathrm{𝟏},0).`$ (5) We here describe the case with one generation of $`L`$ and $`L^c`$ for simplicity. It is straightforward to extend the analysis for several generations of vectorlike leptons, and similar results are obtained for the lepton flavor violating processes. We examine two typical models in the following, for which distinct results will be obtained concerning the supersymmetric contributions to the lepton number violating processes such as $`\mu e\gamma `$. In the first model, the superpotential relevant for the leptonic superfields is given by $`𝒲=\lambda _1^{aj}l_aH_1e_j^c+\lambda _2^{a4}l_aL^cS+\lambda _3H_1H_2S+{\displaystyle \frac{\lambda _4}{3}}S^3,`$ (6) where $`a=i,4`$ and $`i,j=1,2,3`$. The Higgs fields include the pair of $`\mathrm{SU}(2)_\mathrm{L}`$ doublets $`H_1`$ and $`H_2`$ and the singlet $`S`$. The vacuum expectation value (VEV) of $`S`$ gives the mass terms involving the exotic lepton doublets $`L`$ and $`L^c`$ as well as the $`\mu `$ term for the Higgs doublets $`H_1`$ and $`H_2`$. The $`R`$-parity invariance is assumed for the lepton number conservation. The $`\mathrm{Z}_3`$ symmetry is also assumed to exclude the mass term of $`S`$ which is independent of the electroweak scale. (The cosmological domain wall problem associated with the spontaneous breaking of $`\mathrm{Z}_3`$ may be evaded by introducing small explicit breaking in the higher order terms.) The coupling matrices in eq.(6) may be simplified without loss of generality by redefinition of the relevant fields at some unification scale such as the gravitational scale $`M_\mathrm{G}`$. The elements $`\lambda _1^{4j}`$ in the $`4\times 3`$ coupling matrix $`\lambda _1`$ can be eliminated by an appropriate unitary transformation among the left-handed doublets $`l_i`$ and $`l_4L`$. Then, the submatrix $`\lambda _1^{ij}`$ can be diagonalized by a redefinition of the ordinary leptons. Accordingly, we may take $`\lambda _1(M_\mathrm{G})=\left(\begin{array}{c}\delta _{ij}\lambda _1^{ii}\\ \mathrm{𝟎}\end{array}\right).`$ (9) Furthermore, the components of $`\lambda _1\lambda _4`$ can all be made real and non-negative. In the second model, the above form of $`\lambda _1`$ is realized in terms of another $`\mathrm{Z}_3`$, under which the nontrivial transformations of the relevant fields are given by $`L\omega L,L^c\omega ^{}L^c,S\omega S,H_2\omega ^{}H_2`$ (10) with $`\omega ^3=1`$. Then, we have a bare mass term for $`L`$ and $`L^c`$ by making the following replacement in the superpotential (6): $`\lambda _2^{44}LL^cS\text{[model (1)]}M_LLL^c\text{[model (2)]}.`$ (11) In the lepton basis where the $`\lambda _1(M_\mathrm{G})`$ is given by eq.(9), the ordinary-exotic lepton couplings $`\lambda _2^{i4}`$ provide the source of lepton flavor violation. The lepton flavor violating elements in $`\lambda _1`$ are then generated at $`M_W`$ through the renormalization group evolution with these couplings $`\lambda _2^{i4}`$. The renormalization group equations for the Yukawa couplings are given in the model (1) by $`\mu {\displaystyle \frac{d}{d\mu }}\lambda _1^{aj}={\displaystyle \frac{1}{(4\pi )^2}}`$ $`[3{\displaystyle \underset{k=1}{\overset{3}{}}}{\displaystyle \underset{b=1}{\overset{4}{}}}\lambda _1^{ak}\lambda _1^{bk}\lambda _1^{bj}+{\displaystyle \underset{b=1}{\overset{4}{}}}\lambda _1^{bj}\lambda _2^{a4}\lambda _2^{b4}`$ (12) $`+\lambda _1^{aj}\{{\displaystyle \underset{b=1}{\overset{4}{}}}{\displaystyle \underset{k=1}{\overset{3}{}}}(\lambda _1^{bk})^2+\lambda _3^2+3\lambda _4^2\}\lambda _1^{aj}(3g_2^2+3g_1^2)],`$ $`\mu {\displaystyle \frac{d}{d\mu }}\lambda _2^{a4}={\displaystyle \frac{1}{(4\pi )^2}}`$ $`[\lambda _2^{a4}\{4{\displaystyle \underset{b=1}{\overset{4}{}}}(\lambda _2^{b4})^2+2\lambda _3^2+2\lambda _4^2\}`$ (13) $`+{\displaystyle \underset{b=1}{\overset{4}{}}}{\displaystyle \underset{i=1}{\overset{3}{}}}\lambda _2^{b4}\lambda _1^{ai}\lambda _1^{bi}\lambda _2^{a4}(3g_2^2+g_1^2)].`$ The renormalization group equations in the model (2) are also obtained by setting $`\lambda _2^{44}=0`$ in the above equations. Then, we have the leading corrections for the off-diagonal elements of the $`\lambda _1`$ coupling as $`\lambda _1^{aj}(M_W)\lambda _1^{aj}(M_\mathrm{G}){\displaystyle \frac{1}{(4\pi )^2}}\lambda _1^{jj}\lambda _2^{j4}\lambda _2^{a4}\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}}(aj).`$ (14) Hence, even if $`\lambda _1^{aj}(M_\mathrm{G})=0`$ ($`aj`$) at $`M_\mathrm{G}`$ as given in eq.(9), these flavor changing elements of the $`\lambda _1`$ coupling arise at $`M_W`$. It should here be remarked that these corrections for $`\lambda _1`$ are in fact proportional to the corresponding diagonal elements $`\lambda _1^{jj}`$. It should also be mentioned that in the model (2) $`\lambda _1^{4j}0`$ and $`\lambda _2^{44}0`$, which is ensured by the $`\mathrm{Z}_3`$ symmetry (10). The mass terms of the leptons are produced by the vacuum expectation values (VEV’s) of the Higgs fields, $`H_1=\left(\begin{array}{c}v_1\\ 0\end{array}\right),H_2=\left(\begin{array}{c}0\\ v_2\end{array}\right),S=v_3,`$ (19) where $`v=(v_1^2+v_2^2)^{1/2}174\mathrm{G}\mathrm{e}\mathrm{V}`$ and $`\mathrm{tan}\beta =v_2/v_1`$. These VEV’s may acquire nonvanishing phases introducing a new source of $`CP`$ violation. We, however, do not consider this possibility for simplicity, and therefore take the real VEV’s. Then, the charged lepton mass matrix is given at $`M_W`$ by $`M_{}=\left(\begin{array}{cc}\lambda _1^{ij}v_1& \lambda _2^{i4}v_3\\ \lambda _1^{4j}v_1& \lambda _2^{44}v_3\end{array}\right).`$ (22) It should here be remembered that $`\lambda _2^{44}v_3M_L`$ in the model (2). This $`4\times 4`$ real matrix is diagonalized by two orthogonal matrices $`U_{}`$ and $`V_{}`$ as $`U_{}^\mathrm{T}M_{}V_{}=\mathrm{diag}.(m_e,m_\mu ,m_\tau ,M_E).`$ (23) The neutral lepton mass matrix is also produced as $`M_𝒩=\left(\begin{array}{c}\lambda _2^{i4}v_3\\ \lambda _2^{44}v_3\end{array}\right).`$ (26) This $`4\times 1`$ real matrix is diagonalized with an orthogonal matrix $`O_𝒩`$ as $`O_𝒩^\mathrm{T}M_𝒩=\left(\begin{array}{c}\mathrm{𝟎}\\ M_N\end{array}\right).`$ (29) The masses of the ordinary charged leptons are given by $`m_{e_i}\lambda _1^{ii}v_1,`$ (30) where the radiative corrections (14) for the $`\lambda _1`$ coupling are considered. The masses of the exotic leptons are also given by $`M_EM_N=v_3\left[{\displaystyle \underset{a=1}{\overset{4}{}}}(\lambda _2^{a4})^2\right]^{1/2}.`$ (31) The zero modes in eq.(29) are the three generations of ordinary neutrinos. It is noticed that the orthogonal matrix $`O_𝒩`$ is not determined uniquely at this stage. An arbitrary $`3\times 3`$ transformation, say $`O_\nu ^{}`$, to redefine the (approximately) massless ordinary neutrinos can be incorporated in $`O_𝒩`$. The ordinary neutrinos may acquire very small Majorana masses through the seesaw mechanism with the right-handed neutrinos of $`\mathrm{SU}(2)_\mathrm{L}`$ singlet. Then, the neutrino mixing matrix $`O_𝒩`$ is determined completely. We here consider some relevant properties of the transformation matrices $`U_{}`$, $`V_{}`$, and $`O_𝒩`$. First, the orthogonal transformation $`V_{}`$, which represents the right-handed charged lepton mixing, diagonalizes the matrix $`M_{}^\mathrm{T}M_{}`$. By considering the structure of $`M_{}^\mathrm{T}M_{}`$ from eq.(22) with eq.(14), the right-handed ordinary-exotic lepton mixing in $`V_{}`$, which is relevant for describing the gauge interactions as seen below, is estimated in the leading order (with $`M_Ev_3`$ for definiteness) as $`(V_{})_{i4}(V_{})_{4i}(m_{e_i}/M_E)\lambda _2^{i4}.`$ (32) It should here be remarked that the right-handed lepton mixing effects are suppressed sufficiently by the mass ratios $`m_{e_i}/M_E`$ with the experimental bound $`M_E>97.0\mathrm{GeV}`$ l3 , even if the ordinary-exotic lepton couplings $`\lambda _2^{i4}`$ take significant values. On the other hand, the left-handed ordinary-exotic lepton mixings in $`U_{}`$ and $`O_𝒩`$ are estimated as $`(U_{})_{i4}(U_{})_{4i}(O_𝒩)_{i4}(O_𝒩)_{4j}(O_\nu ^{})_{ji}\lambda _2^{i4},`$ (33) where the ambiguity in $`O_𝒩`$ as mentioned above is extracted explicitly by suitably choosing the orthogonal transformation $`O_\nu ^{}`$ of the ordinary neutrinos. These left-handed mixings are no longer suppressed by $`m_{e_i}/M_E`$. Here, we note the similarity between $`U_{}`$ for the charged leptons and $`O_𝒩`$ for the neutral leptons. In fact, the effective combination of left-handed lepton mixing is estimated as $`(U_{}^\mathrm{T}O_𝒩)_{i4}(U_{}^\mathrm{T}O_𝒩)_{4j}(O_\nu ^{})_{ji}(m_{e_i}/M_E)^2\lambda _2^{i4},`$ (34) $`(U_{}^\mathrm{T}O_𝒩)_{ik}(O_\nu ^{})_{kj}\delta _{ij}(m_{e_i}/M_E)(m_{e_j}/M_E)\lambda _2^{i4}\lambda _2^{j4}.`$ (35) This similarity of the left-handed lepton mixings described in terms of the $`\lambda _2^{i4}`$ couplings and the ordinary-exotic lepton mass ratios is indeed stable against the radiative corrections. It can be understood by noting the following facts. (i) By setting $`\lambda _1^{ii}=0`$ at $`M_\mathrm{G}`$ in eq.(14), we have $`\lambda _1^{aj}0`$ at any scales as seen from eq.(12). Then, $`M_{}`$ essentially coincides with $`M_𝒩`$, and they are diagonalized with the left-handed lepton transformation $`U_{}=O_𝒩`$ to eliminate the $`\lambda _2^{i4}v_3`$ term. (ii) The change $`\lambda _1\lambda _1`$ of the Yukawa coupling in $`M_{}`$ is compensated by the change $`e_i^ce_i^c`$ of the right-handed ordinary leptons. This means that the sign change of $`\lambda _1`$ does not affect the left-handed lepton mixing. Hence, by considering (i) and (ii) we find that the difference between $`U_{}`$ and $`O_𝒩`$ should appear at the second order of $`\lambda _1^{aj}(M_W)\lambda _1^{jj}m_{e_j}/v_1`$. (iii) By setting $`\lambda _2^{i4}(M_\mathrm{G})=0`$ so that $`\lambda _2^{i4}=0`$ at any scales, the $`\lambda _1`$ coupling remains to be diagonal even if the radiative corrections are included. Hence, in this limit with $`\lambda _2^{i4}=0`$, $`M_{}`$ and $`M_𝒩`$ are diagonal, and the complete similarity follows with $`U_{}=O_𝒩=\mathrm{𝟏}_4`$ ($`4\times 4`$ unit matrix). It can be shown that the flavor violating gauge couplings of leptons are suppressed naturally due to the smallness of $`(V_{})_{i4}`$ and the similarity between $`U_{}`$ and $`O_𝒩`$. The charged gauge interaction coupled to $`W`$ boson is given by $`_W={\displaystyle \frac{g_2}{\sqrt{2}}}(U_{}^\mathrm{T}O_𝒩)_{ab}\overline{}_a\gamma ^\mu P_\mathrm{L}𝒩_bW_\mu ^{}+{\displaystyle \frac{g_2}{\sqrt{2}}}(V_{})_{4a}\overline{}_a\gamma ^\mu P_\mathrm{R}𝒩_4W_\mu ^{}+\mathrm{h}.\mathrm{c}.,`$ (36) where $`_a(e_i,E)`$ and $`𝒩_a(\nu _i,N)`$ ($`a=14`$) represent the four-component Dirac spinors of the lepton mass eigenstates, and $`P_\mathrm{L}`$ and $`P_\mathrm{R}`$ are the chiral projections to the left-handed and right-handed parts, respectively. This charged gauge interaction is relevant for the $`\mu e\gamma `$ decay. The lepton flavor violating terms with the factors $`(U_{}^\mathrm{T}O_𝒩)_{ab}`$ ($`ab`$) are apparently induced in the left-handed couplings of eq.(36). However, due to the similarity between $`U_{}`$ and $`O_𝒩`$ as shown in eqs.(34) and (35), they are almost rotated out with the suitable redefinition $`O_\nu ^{}`$ of the ordinary neutrinos neglecting the tiny neutrino masses $`m_{\nu _i}`$. The flavor changing right-handed couplings are also suppressed by the small factors $`(V_{})_{4i}`$, as shown in eq.(32). Hence, the effects of the charged gauge interaction appear to be extremely small for the lepton flavor violating processes. The neutral gauge interaction of charged leptons coupled to $`Z`$ boson is given by $`_Z=_Z^0+{\displaystyle \frac{1}{2}}\left(g_2\mathrm{cos}\theta _W+g_1\mathrm{sin}\theta _W\right)(V_{})_{4a}(V_{})_{4b}\overline{}_a\gamma ^\mu P_\mathrm{R}_bZ_\mu ,`$ (37) where $`_Z^0`$ represents the usual flavor diagonal part which is obtained by turning off the ordinary-exotic lepton mixing. The lepton flavor violating term appears only in the right-handed couplings of eq.(37). This is because the left-handed leptons all form the $`\mathrm{SU}(2)_\mathrm{L}`$ doublets. Since the FCNC’s of ordinary leptons in eq.(37) are proportional to $`(V_{})_{4i}(V_{})_{4j}`$, their effects are fairly suppressed by the small mass ratios $`(m_{e_i}/M_E)(m_{e_j}/M_E)`$. If the quark singlets $`(D^c,D)`$ are also present, they may have the flavor violating coupling which has the common $`\mathrm{E}_6`$ origin with the $`\lambda _2`$ coupling of $`(L,L^c)`$. In this case, the FCNC’s of ordinary quarks appear as well to be negligibly small due to the suppression factors $`(m_{d_i}/M_D)(m_{d_j}/M_D)`$. We now estimate the contributions to the $`\mu e\gamma `$ decay which are provided by these gauge interactions. We will in fact observe below that these contributions are negligibly small. The decay amplitude of $`\mu e\gamma `$ is generally given by $`T(\mu e\gamma )=eϵ^\alpha \overline{u}_e(pq)\left[i\sigma _{\alpha \beta }q^\beta (A_\mathrm{L}P_\mathrm{L}+A_\mathrm{R}P_\mathrm{R})\right]u_\mu (p).`$ (38) Then, the decay rate is calculated by $`\mathrm{\Gamma }(\mu e\gamma )={\displaystyle \frac{1}{16\pi }}e^2m_\mu ^3(|A_\mathrm{L}|^2+|A_\mathrm{R}|^2).`$ (39) The contributions to the amplitudes $`A_\mathrm{L}`$ and $`A_\mathrm{R}`$ are provided by the one-loop diagrams with the intermediate states of the neural leptons ($`W`$ mediated) and charged leptons ($`Z`$ mediated). We can see from eqs.(32) – (35) that all these diagrams include the significant suppression factor $`(m_e/M_E)(m_\mu /M_E)10^8`$. Then, the $`W`$ and $`Z`$ contributions to the decay amplitude are given by $`A_{\mathrm{L},\mathrm{R}}^{(W)}A_{\mathrm{L},\mathrm{R}}^{(Z)}{\displaystyle \frac{1}{32\pi ^2}}{\displaystyle \frac{m_em_\mu ^2}{M_W^2M_E^2}}\lambda _2^{14}\lambda _2^{24}.`$ (40) Therefore, the flavor violating gauge couplings induced by the ordinary-exotic lepton mixing provide tiny contributions $`10^{21}`$ to the branching ratio $`\mathrm{Br}(\mu e\gamma )`$. This feature is actually confirmed by numerical calculations. ## III The slepton mixing In the supersymmetric model, significant effects of flavor violation may be obtained through the slepton mixing. It is expected that the supersymmetry breaking is provided in the minimal supergravity model with the soft terms at the gravitational scale $`M_\mathrm{G}10^{18}\mathrm{GeV}`$, including the common slepton masses squared $`m_{\stackrel{~}{l}}^{2}{}_{ab}{}^{}=m_0^2\delta _{ab}`$, etc., the $`A`$ terms $`A_k=a_0m_0\lambda _k`$, and the gaugino masses $`m_1`$ and $`m_2`$. We have seen that in the present sort of models the original source of lepton flavor violation is the ordinary-exotic lepton couplings $`\lambda _2^{i4}`$. At the tree-level with the universal soft supersymmetry breaking terms, however, the flavor violating effects of these couplings are still suppressed by the ordinary-exotic lepton mass ratios $`m_{e_i}/M_E`$, since the lepton and slepton mass matrices are diagonalized simultaneously in this limit. Significant effects of flavor violation are rather provided from the radiative corrections on the soft terms which are generated with the ordinary-exotic lepton couplings $`\lambda _2^{i4}`$. The soft mass terms and $`A`$-terms at the electroweak scale $`M_W`$ deviate from the universal forms through the renormalization group evolution Hall:1986dx . In the leading order approximation the corrections are given for the model (1) by $`\mathrm{\Delta }m_{\stackrel{~}{l}}^{2}{}_{aa}{}^{}`$ $`=`$ $`{\displaystyle \frac{1}{(4\pi )^2}}\left[2m_0^2(3+a_0^2)\left\{(\lambda _1^{aa})^2+(\lambda _2^{a4})^2\right\}(6m_2^2g_2^2+2m_1^2g_1^2)\right]\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}},`$ (41) $`\mathrm{\Delta }m_{\stackrel{~}{l}}^{2}{}_{ab}{}^{}`$ $`=`$ $`{\displaystyle \frac{2}{(4\pi )^2}}m_0^2(3+a_0^2)\lambda _2^{a4}\lambda _2^{b4}\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}}(ab),`$ (42) $`\mathrm{\Delta }m_{\stackrel{~}{e}^c}^{2}{}_{ij}{}^{}`$ $`=`$ $`{\displaystyle \frac{\delta _{ij}}{(4\pi )^2}}\left[4m_0^2(\lambda _1^{ii})^2(3+a_0^2)8g_1^2m_1^2\right]\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}},`$ (43) $`\mathrm{\Delta }m_{\stackrel{~}{L}^c}^2`$ $`=`$ $`{\displaystyle \frac{1}{(4\pi )^2}}\left[2m_0^2(3+a_0^2){\displaystyle \underset{a=1}{\overset{4}{}}}(\lambda _2^{a4})^2(6g_2^2m_2^2+2g_1^2m_1^2)\right]\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}},`$ (44) $`\mathrm{\Delta }A_1^{ii}`$ $`=`$ $`{\displaystyle \frac{1}{(4\pi )^2}}\lambda _1^{ii}[\text{}3a_0m_0\{(\lambda _2^{i4})^2+\lambda _3^2+3\lambda _4^2+3(\lambda _1^{ii})^2+{\displaystyle \underset{j=1}{\overset{3}{}}}(\lambda _1^{jj})^2\}`$ (45) $`3a_0m_0(g_2^2+g_1^2)6(g_2^2m_2+g_1^2m_1)\text{}]\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}},`$ $`\mathrm{\Delta }A_1^{aj}`$ $`=`$ $`{\displaystyle \frac{3}{(4\pi )^2}}a_0m_0\lambda _1^{jj}\lambda _2^{a4}\lambda _2^{j4}\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}}(aj),`$ (46) $`\mathrm{\Delta }A_2^{a4}`$ $`=`$ $`{\displaystyle \frac{1}{(4\pi )^2}}\lambda _2^{a4}[\text{}3a_0m_0\{4{\displaystyle \underset{b=1}{\overset{4}{}}}(\lambda _2^{b4})^2+2\lambda _3^2+2\lambda _4^2+(\lambda _1^{aa})^2\}`$ (47) $`a_0m_0(3g_2^2+g_1^2)2(3g_2^2m_2+g_1^2m_1)\text{}]\mathrm{log}{\displaystyle \frac{M_\mathrm{G}}{M_W}}.`$ We should set $`\lambda _2^{44}=0`$ for the model (2) in these formulas. Because of these radiative corrections, the slepton mass matrices and the corresponding lepton mass matrices are no longer diagonalized simultaneously, providing the new source of lepton flavor violation in the slepton mixing. Then, this slepton mixing is expected to contribute significantly to the $`\mu e\gamma `$ decay. The lepton-slepton interactions with the neutralinos $`\chi _\alpha ^0`$ ($`\alpha =15`$) and the charginos $`\chi _\kappa ^\pm `$ ($`\kappa =1,2`$), which are relevant for $`\mu e\gamma `$, are given, respectively, by $`_{\chi ^0}=N_{a\alpha A}^\mathrm{R}\overline{}_aP_\mathrm{R}\chi _\alpha ^0\stackrel{~}{}_A+N_{a\alpha A}^\mathrm{L}\overline{}_aP_\mathrm{L}\chi _\alpha ^0\stackrel{~}{}_A+\mathrm{h}.\mathrm{c}.`$ (48) and $`_{\chi ^\pm }=C_{a\kappa K}^\mathrm{R}\overline{}_aP_\mathrm{R}\chi _\kappa ^{}\stackrel{~}{𝒩}_K+C_{a\kappa K}^\mathrm{L}\overline{}_aP_\mathrm{L}\chi _\kappa ^{}\stackrel{~}{𝒩}_K+\mathrm{h}.\mathrm{c}.,`$ (49) where $`\stackrel{~}{}_A`$ ($`A=18`$) and $`\stackrel{~}{𝒩}_K`$ ($`K=15`$) are the slepton mass eigenstates. The coupling coefficients are given in terms of the transformation matrices to diagonalize the mass matrices of leptons, sleptons, charginos, and neutralinos. We here evaluate the decay rate of $`\mu e\gamma `$ which is provided by the slepton mixing. By calculating the one-loop diagrams for the $`\mu e\gamma `$ process mediated by the neutralinos and charginos, as shown in fig.1, we obtain the contributions to the decay amplitudes hisano $`A_\mathrm{R}^{(\mathrm{n})}`$ $`=`$ $`{\displaystyle \frac{1}{32\pi ^2}}{\displaystyle \underset{A,\alpha }{}}\left[{\displaystyle \frac{m_\mu }{M_{\stackrel{~}{}_A}^2}}N_{1\alpha A}^\mathrm{R}N_{2\alpha A}^Rf_1(x_{\alpha A})+{\displaystyle \frac{M_{\chi _\alpha ^0}}{M_{\stackrel{~}{}_A}^2}}N_{1\alpha A}^\mathrm{R}N_{2\alpha A}^Lf_2(x_{\alpha A})\right],`$ (50) $`A_\mathrm{L}^{(\mathrm{n})}`$ $`=`$ $`A_\mathrm{R}^{(\mathrm{n})}(RL),`$ (51) $`A_\mathrm{R}^{(\mathrm{c})}`$ $`=`$ $`{\displaystyle \frac{1}{32\pi ^2}}{\displaystyle \underset{K,\kappa }{}}\left[{\displaystyle \frac{m_\mu }{M_{\stackrel{~}{𝒩}_K}^2}}C_{1\kappa K}^\mathrm{R}C_{2\kappa K}^Rf_3(x_{\kappa K})+{\displaystyle \frac{M_{\chi _\kappa ^{}}}{M_{\stackrel{~}{𝒩}_K}^2}}C_{1\kappa K}^\mathrm{R}C_{2\kappa K}^Lf_4(x_{\kappa K})\right],`$ (52) $`A_\mathrm{L}^{(\mathrm{c})}`$ $`=`$ $`A_\mathrm{R}^{(\mathrm{c})}(RL),`$ (53) where $`f_1f_4`$ are certain functions of $`x_{\alpha A}(M_{\chi _\alpha ^0}/M_{\stackrel{~}{}_A})^2`$ and $`x_{\kappa K}(M_{\chi _\kappa ^{}}/M_{\stackrel{~}{𝒩}_K})^2`$. We estimate below $`\mathrm{Br}(\mu e\gamma )`$ with these contributions for the model (1) and (2). We will see that these models provide quite different results. In fact, these supersymmetric contributions are rather suppressed in the model (1) with the extra factor involving $`m_e`$ as well as $`m_\mu `$. ### III.1 Model (1) The charged slepton mass matrix is given as $`_\stackrel{~}{}^2=\left(\begin{array}{cc}M_{\stackrel{~}{}_{\mathrm{LL}}}^2& M_{\stackrel{~}{}_{\mathrm{LR}}}^2\\ M_{\stackrel{~}{}_{\mathrm{RL}}}^2& M_{\stackrel{~}{}_{\mathrm{RR}}}^2\end{array}\right),`$ (56) where $`\mathrm{L}`$ and $`\mathrm{R}`$ represent the chirality of the corresponding leptons. The neutral slepton mass matrix is also given in a similar form. In order to see the flavor changing structure of the lepton-slepton-gaugino interactions, it is suitable to transform this slepton mass matrix by the orthogonal matrices which diagonalize the lepton mass matrix: $`_{\stackrel{~}{}}^{2}{}_{}{}^{}=\left(\begin{array}{cc}U_{}^\mathrm{T}& 0\\ 0& V_{}^\mathrm{T}\end{array}\right)_\stackrel{~}{}^2\left(\begin{array}{cc}U_{}& 0\\ 0& V_{}\end{array}\right).`$ (61) The universality for the soft supersymmetry breaking at $`M_\mathrm{G}`$ is violated at $`M_W`$ due to the renormalization group effects. Then, the reduced slepton mass matrix $`_{\stackrel{~}{}}^{2}{}_{}{}^{}`$ involves the flavor changing elements. In the model (1), the components of the slepton mass matrix are given including the renormalization group effects as $`M_{\stackrel{~}{}_{\mathrm{LL}}}^2`$ $`=`$ $`x_{\mathrm{LL}}M_{}M_{}^\mathrm{T}+y_{\mathrm{LL}}m_0^2\mathrm{𝟏}_4+m_0^2\left(\begin{array}{cc}c_{\mathrm{LL}}^{ij}(\lambda _1^{ii})^2+c_{\mathrm{LL}}^{ji}(\lambda _1^{jj})^2& d_{\mathrm{LL}}^i(\lambda _1^{ii})^2\\ d_{\mathrm{LL}}^j(\lambda _1^{jj})^2& 0\end{array}\right),`$ (64) $`M_{\stackrel{~}{}_{\mathrm{LR}}}^2`$ $`=`$ $`(M_{\stackrel{~}{}_{\mathrm{RL}}}^2)^\mathrm{T}`$ (68) $`=`$ $`x_{\mathrm{LR}}m_0M_{}+x_{\mathrm{LR}}^{}m_0\left(\begin{array}{cc}\mathrm{𝟎}& M_𝒩\end{array}\right)+m_0^2\left(\begin{array}{cc}\delta _{ij}f_{\mathrm{LR}}^i\lambda _1^{ii}& g_{\mathrm{LR}}^i(\lambda _1^{ii})^2\\ \mathrm{𝟎}& 0\end{array}\right),`$ $`M_{\stackrel{~}{}_{\mathrm{RR}}}^2`$ $`=`$ $`x_{\mathrm{RR}}M_{}^\mathrm{T}M_{}+y_{\mathrm{RR}}m_0^2\mathrm{𝟏}_4+z_{\mathrm{RR}}m_0^2\left(\begin{array}{cc}\mathrm{𝟎}& \mathrm{𝟎}\\ \mathrm{𝟎}& 1\end{array}\right)+m_0^2\left(\begin{array}{cc}\delta _{ij}c_{\mathrm{RR}}^i(\lambda _1^{ii})^2& \mathrm{𝟎}\\ \mathrm{𝟎}& 0\end{array}\right),`$ (73) where $`x_{\mathrm{LL}},\mathrm{}1`$ are the relevant parameters. The terms proportional to the $`4\times 4`$ unit matrix $`\mathrm{𝟏}_4`$ as well as those given by the charged lepton mass matrix $`M_{}`$ are diagonalized by the transformation in eq.(61). The term given by the neutral lepton mass matrix $`M_𝒩`$ is almost diagonalized by the similarity of $`U_{}^\mathrm{T}O_𝒩`$. The third term of eq.(73) provides a contribution $`(V_{})_{4i}(V_{})_{4j}`$ for the slepton mixing, which is substantially suppressed by the lepton mass ratios $`(m_{e_i}/M_E)(m_{e_j}/M_E)`$. The remaining terms involve the Yukawa couplings $`\lambda _1^{ii}m_{e_i}/v_1`$. Among these contributions in the model (1), the significant flavor mixing arises from the fourth term of eq.(73) in the right-handed charged slepton sector as $`(M_{\stackrel{~}{}_{\mathrm{RR}}}^2)_{ij}/m_0^2(V_{}^\mathrm{T})_{ii}(\lambda _1^{ii})^2(V_{})_{ij}+(V_{}^\mathrm{T})_{ij}(\lambda _1^{jj})^2(V_{})_{jj}.`$ (74) It is here noted that the right-handed ordinary charged lepton mixing induced by the ordinary-exotic lepton mixing is related to the charged lepton masses as $`(V_{})_{ij}{\displaystyle \frac{m_{e_i}m_{e_j}}{m_{e_i}^2+m_{e_j}^2}}\lambda _2^{i4}\lambda _2^{j4}(ij).`$ (75) By considering these arguments on the slepton mixing, the dominant contribution from the neutralino couplings to the decay amplitude of $`\mu e\gamma `$ is estimated as $`A_\mathrm{L}^{(\mathrm{n})}{\displaystyle \frac{1}{32\pi ^2}}{\displaystyle \frac{\lambda _3\lambda _1^{22}v_2v_3}{m_0^3}}(\lambda _1^{22})^2(m_e/m_\mu )\lambda _2^{14}\lambda _2^{24}.`$ (76) Here, the right-handed $`\stackrel{~}{\mu }^c`$-$`\stackrel{~}{e}^c`$ slepton mixing $`(\lambda _1^{22})^2(V_{})_{12}`$ given by eqs.(74) and (75) is used for the $`e_\mathrm{R}`$-$`\stackrel{~}{\mu }^c`$-$`\chi ^0`$ vertex, and the chirality flip $`\lambda _3\lambda _1^{22}v_2v_3`$ of the intermediate $`\stackrel{~}{\mu }`$ state is provided by the $`|F_{H_1}|^2`$ term. The chargino couplings provide a contribution of the same order as eq.(76), where the chirality flipping $`\stackrel{~}{N}`$-$`\stackrel{~}{e}^c`$ slepton mixing given by $`(O_{}^\mathrm{T}M_{\stackrel{~}{𝒩}_{\mathrm{LR}}}^2)_{41}/m_0(m_e/m_0)\mathrm{tan}\beta \lambda _2^{14}`$ is used for the $`e_\mathrm{R}`$-$`\stackrel{~}{N}`$-$`\chi _0`$ vertex, and the the left-handed $`\stackrel{~}{N}`$-$`\stackrel{~}{\nu }_\mu `$ slepton mixing in the intermediate state is provided by $`(O_{}^\mathrm{T}M_{\stackrel{~}{𝒩}_{\mathrm{LL}}}^2)_{42}/m_0^2(m_\mu /M_W)^2\mathrm{tan}\beta \lambda _2^{24}`$. Accordingly, in the model (1) the contributions to the decay amplitude are dominantly given by $`A_\mathrm{L}^{(\mathrm{n})}A_\mathrm{L}^{(\mathrm{c})}{\displaystyle \frac{1}{32\pi ^2}}\mathrm{tan}^3\beta {\displaystyle \frac{m_em_\mu ^2}{M_W^2m_0^2}}\lambda _2^{14}\lambda _2^{24},`$ (77) where $`v_3m_0`$ is considered for definiteness. These contributions should be compared to the non-supersymmetric contributions given in eq.(40). Although these supersymmetric contributions have the specific dependence on the relevant lepton masses similar to eq.(40), it can be enhanced significantly by the power of $`\mathrm{tan}\beta `$. We roughly estimate the branching ratio as $`\mathrm{Br}(\mu e\gamma )`$ $``$ $`{\displaystyle \frac{3\alpha }{32\pi }}\mathrm{tan}^6\beta \left({\displaystyle \frac{m_em_\mu }{m_0^2}}\right)^2\left(\lambda _2^{14}\lambda _2^{24}\right)^2`$ (78) $``$ $`10^{13}(\mathrm{tan}\beta 20,\lambda _2^{14}\lambda _2^{24}0.3).`$ It should here be remarked as seen in eq.(77) that in the model (1) the chirality of the charged leptons is mainly specified as $`\mu _\mathrm{L}^{}e_\mathrm{R}^{}\gamma `$. The dependence of the decay amplitudes on both $`m_e`$ and $`m_\mu `$, as seen in eq.(77), can be understood as follows by means of approximate flavor symmetries. Set $`\lambda _1^{11}(M_\mathrm{G})=0`$ in eq.(9). Then, we can make $`\lambda _2^{14}(M_\mathrm{G})=0`$ by a suitable transformation of the left-handed lepton doublets $`l_1`$ and $`l_4=L`$, while keeping $`\lambda _1(M_\mathrm{G})`$ with $`\lambda _1^{11}(M_\mathrm{G})=0`$. In this limit, a flavor symmetry $`\mathrm{U}(1)_e`$ appears under the phase transformation $`l_1\mathrm{e}^{i\alpha }l_1`$ and $`e_1^c\mathrm{e}^{i\alpha }e_1^c`$. If the universal form is assumed for the supersymmetry breaking terms at $`M_\mathrm{G}`$, this flavor symmetry $`\mathrm{U}(1)_e`$ is still preserved in the limit $`\lambda _1^{11}(M_\mathrm{G})=0`$. This is the case even if the renormalization group corrections are included. Therefore, we find that in the limit $`\lambda _1^{11}(M_\mathrm{G})=0`$ corresponding to $`m_e=0`$, the $`\mu e\gamma `$ decay is prevented for the electron number conservation by the $`\mathrm{U}(1)_e`$. Similarly, the decay amplitudes of $`\mu e\gamma `$ depend on $`m_\mu `$ as well due to the approximate $`\mathrm{U}(1)_\mu `$ symmetry. It should further be remarked that even for the case with several pairs of vectorlike lepton doublets, the $`\mu e\gamma `$ decay amplitudes are dominantly given by eq.(77) in the model (1) without the bare mass term $`M_LLL^c`$. This can be justified by extending readily the above symmetry argument. In order to confirm the estimate of $`\mathrm{Br}(\mu e\gamma )`$ given in eq.(78), we have made numerical calculations by taking reasonable values of the model parameters. In the model (1), the most important parameters are $`\mathrm{tan}\beta `$ and $`\lambda _2^{i4}`$ for the ordinary-exotic lepton mixing. The values of $`\lambda _1^{ii}`$ are determined so as to reproduce the actual lepton masses $`m_e,m_\mu ,m_\tau `$. The exotic leptons $`E`$ and $`N`$ have typically the masses $`M_EM_N100\mathrm{G}\mathrm{e}\mathrm{V}`$ by taking $`\lambda _2^{44}0.3`$ and $`v_3500\mathrm{G}\mathrm{e}\mathrm{V}`$. In fig.2, Br($`\mu e\gamma `$) is shown as a function of $`\lambda _2^{14}\lambda _2^{24}`$ by taking randomly the values of $`\lambda _2^{i4}`$. We have also taken typically $`\mathrm{tan}\beta =20`$, $`v_3=500\mathrm{G}\mathrm{e}\mathrm{V}`$, $`m_0=100\mathrm{G}\mathrm{e}\mathrm{V}`$, $`a_0=1`$, $`m_1=100\mathrm{G}\mathrm{e}\mathrm{V}`$, $`\lambda _2^{44}=0.3`$, $`\lambda _3=0.6`$ and $`\lambda _4=0.5`$. It is clearly observed in fig.2 that Br($`\mu e\gamma `$) is roughly proportional to $`(\lambda _2^{14}\lambda _2^{24})^2`$, as indicated in eq.(78). The relation between Br($`\mu e\gamma `$) and $`\mathrm{tan}\beta `$ is shown in fig.3. Here, we have taken typically $`v_3=500\mathrm{G}\mathrm{e}\mathrm{V}`$, $`m_0=200\mathrm{G}\mathrm{e}\mathrm{V}`$, $`a_0=0`$, $`m_1=200\mathrm{G}\mathrm{e}\mathrm{V}`$, $`\lambda _2^{a4}=0.3`$, $`\lambda _3=0.6`$ and $`\lambda _4=0.5`$. This result indicates the strong $`\mathrm{tan}\beta `$ dependence as $`\mathrm{Br}(\mu e\gamma )\mathrm{tan}^6\beta `$. It is, in particular, interesting that if $`\mathrm{tan}\beta `$ is larger than 20, $`\mathrm{Br}(\mu e\gamma )`$ might be comparable to the experimental bound $`1.2\times 10^{11}`$ brbound . As for the $`\mu 3e`$ decay, the penguin diagrams associated with fig.1 provide dominant contributions in the present model with $`\mathrm{Br}(\mu 3e)/\mathrm{Br}(\mu e\gamma )7\times 10^3`$, as is usually the case hisano . The tree-level $`Z`$ mediated FCNC’s given in eq.(37) provide only a contribution smaller by a few orders. We have also the supersymmetric contributions to $`\tau e\gamma `$ and $`\tau \mu \gamma `$, which are estimated in eq.(78) by replacing the relevant charged lepton masses and the $`\lambda _2^{i4}`$ couplings. In particular, $`\mathrm{Br}(\tau \mu \gamma )`$ is enhanced as $`\mathrm{Br}(\tau \mu \gamma )`$ $``$ $`\left({\displaystyle \frac{m_\mu m_\tau }{m_em_\mu }}\right)^2\mathrm{Br}(\mu e\gamma )`$ (79) $``$ $`10^7(\mathrm{tan}\beta 20,\lambda _2^{24}\lambda _2^{34}0.3),`$ which can be comparable to the experimental bound $`1.1\times 10^6`$ Ahmed:1999gh . Therefore, in the model (1) the $`\tau \mu \gamma `$ decay seems to be more promising to observe the lepton flavor violation due to the ordinary-exotic lepton mixing with supersymmetry. The $`\lambda _2`$ coupling may also contribute to the flavor violation in the quark sector if the quark singlets $`(D^c,D)`$ survive as well to the electroweak scale in the $`\mathrm{E}_6`$ unification. Then, we have dominantly the right-handed squark mixings such as $`(M_{\stackrel{~}{𝒟}_{\mathrm{RR}}}^2)_{12}/m_0^2(m_sm_d/v^2)\lambda _2^{14}\lambda _2^{24}\mathrm{tan}^2\beta 10^8\lambda _2^{14}\lambda _2^{24}\mathrm{tan}^2\beta `$, which are the same as the right-handed slepton mixings given in eqs.(74) with (75). It is clearly found that even for $`\mathrm{tan}\beta 50`$ this squark mixing is much smaller than the bound $`10^2`$ obtained from the $`K^0`$-$`\overline{K}^0`$ system Gabbiani:1996 . ### III.2 Model (2) In the model (2), we have more significant contributions for the right-handed decay amplitudes. This is because the $`\mathrm{U}(1)_e\times \mathrm{U}(1)_\mu `$ considered in the case of model (1) is substantially violated due to the presence of bare mass term $`M_LLL^c`$ of the vectorlike lepton doublets. Actually, the soft supersymmetry breaking mass terms for the ordinary slepton doublets $`\stackrel{~}{l}`$ and the vectorlike slepton doublet $`\stackrel{~}{L}`$ acquire the renormalization group corrections in somewhat different way with $`\lambda _2^{44}0`$. Then, we have an extra terms for $`M_{\stackrel{~}{}_{\mathrm{LL}}}^2`$ and $`M_{\stackrel{~}{𝒩}_{\mathrm{LL}}}^2`$ as $`z_{\mathrm{LL}}m_0^2\left(\begin{array}{cc}\mathrm{𝟎}& \mathrm{𝟎}\\ \mathrm{𝟎}& 1\end{array}\right)M_{\stackrel{~}{}_{\mathrm{LL}}}^2,M_{\stackrel{~}{𝒩}_{\mathrm{LL}}}^2.`$ (82) This extra term provides the significant mixing for the left-handed sleptons as $`(M_{\stackrel{~}{}_{\mathrm{LL}}}^2)_{ij}/m_0^2(M_{\stackrel{~}{𝒩}_{\mathrm{LL}}}^2)_{ij}/m_0^2(U_{}^\mathrm{T})_{i4}(U_{})_{4j}\lambda _2^{i4}\lambda _2^{j4},`$ (83) where the similarity $`U_{}O_𝒩`$ is considered. These left-handed slepton mixings can be taken for the one-loop diagrams similar to those considered in the model (1) with chirality change $`\mathrm{L}\mathrm{R}`$. Then, we obtain the contributions to the $`\mu e\gamma `$ decay amplitude as $`A_\mathrm{R}^{(\mathrm{n})}A_\mathrm{R}^{(\mathrm{c})}{\displaystyle \frac{1}{32\pi ^2}}\mathrm{tan}\beta {\displaystyle \frac{m_\mu }{m_0^2}}\lambda _2^{14}\lambda _2^{24}.`$ (84) It is here important that these supersymmetric contributions are no longer suppressed by the very small extra factor $`m_em_\mu /M_W^2`$ as seen in eq.(77) for the model (1). We roughly estimate the branching ratio as $`\mathrm{Br}(\mu e\gamma )`$ $``$ $`{\displaystyle \frac{3\alpha }{32\pi }}\mathrm{tan}^2\beta \left({\displaystyle \frac{M_W}{m_0}}\right)^4\left(\lambda _2^{14}\lambda _2^{24}\right)^2`$ (85) $``$ $`10^{11}(\mathrm{tan}\beta 1,\lambda _2^{14}\lambda _2^{24}0.01).`$ As seen in eq.(84), the chirality of the charged leptons is mainly specified as $`\mu _\mathrm{R}^{}e_\mathrm{L}^{}\gamma `$ in the model (2) contrary to the case of model (1). It is also mentioned that this significant supersymmetric contribution to $`\mathrm{Br}(\mu e\gamma )`$ may be obtained even in the model (1), if the soft slepton masses squared $`m_{\stackrel{~}{l}}^2`$ are different between $`\stackrel{~}{l}_i`$ and $`\stackrel{~}{l}_4=\stackrel{~}{L}`$ already at $`M_\mathrm{G}`$ for some reason. We have also made numerical calculations to evaluate $`\mathrm{Br}(\mu e\gamma )`$ in the model (2) with the bare mass term, which is expected to be much larger than in the model (1). In fig.4, Br($`\mu e\gamma `$) is shown as a function of $`\lambda _2^{14}\lambda _2^{24}`$ by taking randomly the values of $`\lambda _2^{i4}`$. We have also taken typically $`\mathrm{tan}\beta =5`$, $`v_3=500\mathrm{G}\mathrm{e}\mathrm{V}`$, $`m_0=300\mathrm{G}\mathrm{e}\mathrm{V}`$, $`a_0=1`$, $`m_1=300\mathrm{G}\mathrm{e}\mathrm{V}`$, $`M_L=100\mathrm{G}\mathrm{e}\mathrm{V}`$, $`\lambda _3=0.8`$ and $`\lambda _4=0.7`$. It is clearly observed in fig.4 that Br($`\mu e\gamma `$) is roughly proportional to $`(\lambda _2^{14}\lambda _2^{24})^2`$, as indicated in eq.(85). It is interesting to observe in fig.4 that $`\mathrm{Br}(\mu e\gamma )`$ becomes comparable to the experimental bound $`1.2\times 10^{11}`$ brbound for the rather small ordinary-exotic couplings $`(\lambda _2^{14}\lambda _2^{24})^{1/2}0.02`$ with the moderate value of $`\mathrm{tan}\beta 5`$. The relation between Br($`\mu e\gamma `$) and the gravitino mass $`m_0`$ is shown in fig.5 for $`m_03\mathrm{T}\mathrm{e}\mathrm{V}`$ with somewhat larger $`\lambda _2^{i4}=0.07`$, where the other parameters are taken typically as $`\mathrm{tan}\beta =5`$, $`v_3=500\mathrm{G}\mathrm{e}\mathrm{V}`$, $`a_0=1`$, $`m_1=1000\mathrm{G}\mathrm{e}\mathrm{V}`$, $`M_L=100\mathrm{G}\mathrm{e}\mathrm{V}`$, $`\lambda _3=0.8`$ and $`\lambda _4=0.7`$. It is here noticed that cancellation between the neutralino and chargino contributions occurs for certain values of $`m_0`$, where Br($`\mu e\gamma `$) becomes substantially small. It is also remarkable in the model (2) that even for the relatively large soft supersymmetry breaking with $`m_01\mathrm{TeV}`$, Br($`\mu e\gamma `$) can be comparable to the experimental bound by taking the relatively large ordinary-exotic lepton couplings $`\lambda _2^{i4}0.1`$. We have also the supersymmetric contributions to $`\tau e\gamma `$ and $`\tau \mu \gamma `$, which are estimated in eq.(84) by replacing $`m_\mu `$ with $`m_\tau `$ and the $`\lambda _2^{i4}`$ couplings. The resultant branching ratios of $`\tau e\gamma `$ and $`\tau \mu \gamma `$ are similar to that of $`\mu e\gamma `$, since they are almost independent of $`m_\tau `$ in this case. Therefore, by considering the experimental bounds on the branching ratios, the $`\mu e\gamma `$ seems to be more promising to observe the lepton flavor violation in the model (2) with the bare mass term, if the ordinary-exotic lepton couplings $`\lambda _2^{i4}`$ are comparable each other. The possible contributions of the $`\lambda _2`$ coupling to the flavor violation in the quark sector are estimated in this model (2) as follows. We have here the significant left-handed squark mixings such as $`(M_{\stackrel{~}{𝒟}_{\mathrm{LL}}}^2)_{12}/m_0^2\lambda _2^{14}\lambda _2^{24}`$, which is the same as the left-handed slepton mixings given in eq.(83). Then, a constraint $`\lambda _2^{14}\lambda _2^{24}10^2`$ is placed from the $`K^0`$-$`\overline{K}^0`$ system Gabbiani:1996 . This constraint is, however, less stringent than that from the $`\mu e\gamma `$. ## IV Summary In summary, we have investigated the lepton flavor violating processes, especially the $`\mu e\gamma `$ decay, which are obtained from the mixing between the ordinary leptons and the vectorlike leptons. Although the lepton FCNC’s appear at the tree level, their effects are small naturally because of the hierarchy of the charged lepton masses. Hence, the fine tuning of parameters is not necessary to suppress the FCNC’s sufficiently. In the supersymmetric model, significant effects of lepton flavor violation are obtained from the slepton mixing which is generated by radiative corrections with the ordinary-exotic lepton couplings. In the model (1) without the bare mass term of the vectorlike lepton doublets, the supersymmetric contributions to the $`\mu e\gamma `$ decay are rather suppressed by $`m_e`$ as well as $`m_\mu `$ due to the approximate $`\mathrm{U}(1)_e\times \mathrm{U}(1)_\mu `$ symmetry, which is similar to the gauge boson mediated contributions. It is, however, remarkable that these supersymmetric contributions to $`\mathrm{Br}(\mu e\gamma )`$ are substantially enhanced by the factor $`\mathrm{tan}^6\beta `$ compared to the gauge boson mediated contributions. Then, if $`\mathrm{tan}\beta `$ is larger than 20, $`\mathrm{Br}(\mu e\gamma )`$ might be comparable to the present experimental bound. In the model (2) with the bare mass term, much larger contributions to $`\mathrm{Br}(\mu e\gamma )`$ are obtained through the slepton mixing. In either case we have shown that the slepton mixing contributions to the $`\mu e\gamma `$ decay can be large enough to be observed in the near future experiments. It should also be noted that in the model (1) the $`\tau \mu \gamma `$ decay may be more promising as seen in eq.(79). The discovery of $`\mu e\gamma `$ clearly indicates the new physics. The supersymmetric contributions to the $`\mu e\gamma `$ decay have been considered so far in the literature for other supersymmetric models such as the SUSY GUT Barbieri:1994pv ; Barbieri:1995tw ; hisano2 and the MSSM with right-handed neutrinos Borzumati:1986qx ; hisano ; Hisano:1998fj . The predictions of Br($`\mu e\gamma `$) in these models are within reach of the future experiments. These are, however, the effects of ultra high energy physics much above the electroweak scale. In the supersymmetric model with vectorlike leptons, which may be motivated by the $`\mathrm{E}_6`$ type unification, the significant contributions to the $`\mu e\gamma `$ decay can be obtained due to the exotic leptons and their scalar partners which exist just around the electroweak scale. Therefore, the direct experimental search is feasible for these new particles l3 ; search . It is also observed in the model (2) with the bare mass term of vectorlike leptons that even if the soft supersymmetry breaking scale is around $`1\mathrm{TeV}`$, Br($`\mu e\gamma `$) can be comparable to the current experimental bound with relatively large ordinary-exotic lepton couplings. These features are salient to the supersymmetric model with vectorlike leptons in contrast to the other supersymmetric models. ###### Acknowledgements. We would like to thank Y. Okada and J. Hisano for useful discussions.
warning/0003/hep-th0003297.html
ar5iv
text
# Quantization of Constrained Systems11footnote 1To appear in the proceedings of the 39th Schladming Winter School on “Methods of Quantization”, February 26-March 4, 2000, Schladming, Austria. ## 1 INTRODUCTION ### 1.1 Initial comments The quantization of systems with constraints is important conceptually as well as practically. Principal techniques for the quantization of such systems involve conventional operator techniques , path integral techniques in terms of the original phase space variables , extended operator techniques involving ghost variables in addition to the original variables and extended path integral techniques also including ghost fields (see, e.g., ). However, these standard approaches are generally not unambiguous and may exhibit certain difficulties in application. A recent review carefully analyzes these traditional methods and details their weaknesses as well as their strengths. Canonical quantization generally requires the use of Cartesian coordinates and not more general coordinates . Therefore, whenever we consider a dynamical system without any constraints whatsoever, we assume that the phase space of the unconstrained system is flat and admits a standard quantization of its canonical variables either in an operator form or in an equivalent path integral form. Next, suppose constraints exist, which, for the sake of discussion, we choose as a closed set of first-class constraints; extensions to treat more general constraints are presented in later sections. Whenever there are constraints the original set of variables is no longer composed solely of physical variables but now contains some unphysical variables as well. While such variables cause little concern from a classical standpoint, they are viewed as highly unwelcome from a quantum standpoint inasmuch as one generally wants to quantize only physical variables. Thus it is often deemed necessary to eliminate the unphysical variables leaving only the true physical degrees of freedom. Quantization of the true degrees of freedom is supposed to proceed as in the initial step. In the general case, however, a quantization of the remaining degrees of freedom is not straightforward or perhaps not even possible because the physical (reduced) phase space is non-Euclidean meaning that an obstruction has arisen where none existed before. An obstruction generally precludes the existence of self-adjoint (observable!) canonical operators satisfying the canonical commutation relations. In path integral treatments, such obstructions arise from the introduction of delta functionals that enforce the classical constraints and the concomitant need to introduce subsidiary delta functionals to select a compatible dynamical gauge in order to introduce a canonical sympletic structure on the physical phase space that generally is not flat. These are fundamental problems that seem difficult to overcome. This article reviews a middle ground in the quantization procedure of systems with constraints which may be called the projection-operator, coherent-state approach. Briefly stated, quantization of the original, unconstrained variables proceeds without obstruction or ambiguity, while constraints are enforced by means of a well-chosen projection operator projecting the original Hilbert space onto the physical Hilbert subspace. This conservative framework is presented in the form of a phase-space path integral with the help of coherent states (which, while convenient, are not necessary). The difference between the present approach and other functional integral methods may be attributed to an alternative choice for the integration measure for the Lagrange multiplier variables. The present approach may be traced from . In addition, some aspects of the projection operator approach have been presented in unpublished work of Shabanov ; see also . ### 1.2 Classical backround For our initial discussion, let us briefly review the classical theory of constraints. Let $`\{p_j,q^j\}`$, $`1jJ`$, denote a set of dynamical variables, $`\{\lambda ^a\}`$, $`1aA`$, a set of Lagrange multipliers, and $`\{\varphi _a(p,q)\}`$ a set of constraints. Then the dynamics of a constrained system may be summarized in the form of an action principle by means of the classical action (summation implied) $`I=[p_j\dot{q}^jH(p,q)\lambda ^a\varphi _a(p,q)]𝑑t.`$ (1) The resultant equations that arise from the action read $`\dot{q}^j={\displaystyle \frac{H(p,q)}{p_j}}+\lambda ^a{\displaystyle \frac{\varphi _a(p,q)}{p_j}}\{q^j,H\}+\lambda ^a\{q^j,\varphi _a\},`$ $`\dot{p}_j={\displaystyle \frac{H(p,q)}{q^j}}\lambda ^a{\displaystyle \frac{\varphi _a(p,q)}{q^j}}\{p_j,H\}+\lambda ^a\{p_j,\varphi _a\},`$ $`\varphi _a(p,q)=0,`$ (2) where $`\{,\}`$ denotes the Poisson bracket. The set of conditions $`\{\varphi _a(p,q)=0\}`$ defines the constraint hypersurface. If the constraints satisfy $`\{\varphi _a(p,q),\varphi _b(p,q)\}=c_{ab}^c\varphi _c(p,q),`$ (3) $`\{\varphi _a(p,q),H(p,q)\}=h_a^b\varphi _b(p,q),`$ (4) then we are dealing with a system of first-class constraints. If the coefficients $`c_{ab}^c`$ and $`h_a^b`$ are constants, then it is a closed system of first-class constraints; if they are suitable functions of the variables $`p,q`$, then it is called an open first-class constraint system. If (3) fails, or (3) and (4) fail, then the constraints are said to be second class (see below). For first-class constraints it is sufficient to impose the constraints at the initial time inasmuch as the equations of motion will ensure that the constraints are fulfilled at all future times. Such an initial imposition of the constraints is called an initial value equation. Furthermore, the Lagrange multipliers are not determined by the equations of motion; rather the solutions of the equations of motion depend on them. By specifying the Lagrange multipliers, the solution can be forced to satisfy an additional (“gauge”) condition. Observable quantities are gauge invariant and, hence, do not depend on the gauge abritrariness. For second-class constraints, on the other hand, the Lagrange multipliers are determined by the equations of motion in such a way that the constraints are satisfied for all time. In the remainder of this section we review standard quantization procedures for systems with closed first-class constraints, both of the operator and path integral variety, pointing out some problems in each approach. ### 1.3 Quantization first: Standard <br>operator quantization For a system of closed first-class constraints we assume (with $`\mathrm{}=1`$) that $`[\mathrm{\Phi }_a(P,Q),\mathrm{\Phi }_b(P,Q)]=ic_{ab}^c\mathrm{\Phi }_c(P,Q),`$ (5) $`[\mathrm{\Phi }_a(P,Q),(P,Q)]=ih_a^b\mathrm{\Phi }_b(P,Q),`$ (6) where $`\mathrm{\Phi }_a`$ and $``$ denote self-adjoint constraint and Hamiltonian operators, respectively. Following Dirac , we adopt the quantization prescription given by $`i\dot{W}(P,Q)=[W(P,Q),(P,Q)]`$ (7) where $`W`$ denotes a general function of the kinematical operators $`\{Q^j\}`$ and $`\{P_j\}`$ which are taken as a self-adjoint, irreducible representation of the commutation rules $`[Q^j,P_k]=i\delta _k^j11`$, with all other commutators vanishing. The equations of motion hold for all time $`t`$, say $`0<t<T`$. On the other hand, the conditions $`\mathrm{\Phi }_a(P,Q)|\psi _{phys}=0`$ (8) to select the physical Hilbert space are imposed only at time $`t=0`$ as the analog of the initial value equation; the quantum equations of motion ensure that the constraint conditions are fulfilled for all time. The procedure of Dirac has potential difficulties if zero lies in the continuous spectrum of the constraint operators for in that case there are no normalizable solutions of the constraint condition. We face the same problem, of course, and our resolution is discussed below. ### 1.4 Reduction first: Standard <br>path integral quantization Faddeev has given a path integral formulation in the case of closed first-class constraint systems as follows. The formal path integral $`{\displaystyle \mathrm{exp}\{i_0^T[p_j\dot{q}^jH(p,q)\lambda ^a\varphi _a(p,q)]𝑑t\}𝒟p𝒟q𝒟\lambda }`$ $`={\displaystyle \mathrm{exp}\{i_0^T[p_j\dot{q}^jH(p,q)]𝑑t\}\delta \{\varphi (p,q)\}𝒟p𝒟q}`$ (9) may well encounter divergences in the remaining integrals. Therefore, subsidiary conditions in the form $`\chi ^a(p,q)=0`$, $`1aA`$, are imposed picking out (ideally) one gauge equivalent point per gauge orbit, and in addition a factor in the form of the Faddeev-Popov determinant is introduced to formally preserve canonical covariance. The result is the path integral $`{\displaystyle \mathrm{exp}\{i_0^T[p_j\dot{q}^jH(p,q)]𝑑t\}\delta \{\chi (p,q)\}det(\{\chi ^a,\varphi _b\})\delta \{\varphi (p,q)\}𝒟p𝒟q}.`$ (10) This result may also be expressed as $`{\displaystyle \mathrm{exp}\{i_0^T[p_j^{}\dot{q}^jH^{}(p^{},q^{})]𝑑t\}𝒟p^{}𝒟q^{}},`$ (11) namely, as a path integral over a reduced phase space in which the $`\delta `$-functionals have been used to eliminate $`2A`$ integration variables. The final expression generally involves an integral over a non-Euclidean phase space for which the conventional definition of the path integral is typically ill defined. Thus this widely used prescription is not without its difficulties. ### 1.5 Quantization first $``$ reduction first The two schemes illustrated in the preceding sections are different in principle. In the initial case, one quantizes first and reduces second; in the latter case, one reduces first and quantizes second. For certain systems the results of these different procedures are the same, but that is not universally the case, as we now proceed to illustrate. Let us consider the example of a single degree of freedom specified by the classical action $`I=[p\dot{q}\lambda (p^2+q^4E)]𝑑t.`$ (12) Observe that the classical Hamiltonian vanishes and there is a single constraint. The question we pose is: For what values of $`E`$, $`E>0`$, is the quantum theory nontrivial? On the one hand, according to the procedure of Dirac, the physical Hilbert space is either empty or one-dimensional, spanned by the nonvanishing eigenvector $`|\psi _n`$ that satisfies $`(P^2+Q^4)|\psi _n=E_n|\psi _n,`$ (13) for $`E_n`$ one of the purely discrete eigenvalues for the “Hamiltonian” $`P^2+Q^4`$. On the other hand, the procedure of Faddeev leads initially to $`{\displaystyle e^{i{\scriptscriptstyle p𝑑q}}\delta \{p^2+q^4E\}𝒟p𝒟q}.`$ (14) Next, we fix a gauge, e.g., $`p=0`$, in which case the reduced phase space propagator is given by $`{\displaystyle e^{i{\scriptscriptstyle p𝑑q}}\delta \{p^2+q^4E\}\mathrm{\Pi }(4q^3)\delta \{p\}𝒟p𝒟q}`$ $`=0,`$ (15) which vanishes due to cancellation between the term with $`q>0`$ and the term with $`q<0`$. Note that the symbol $`\mathrm{\Pi }`$ denotes a formal multiplication over all time points. An alternative evaluation may be given if we allow only the term with $`q>0`$, which is achieved by instead using $`{\displaystyle e^{i{\scriptscriptstyle p𝑑q}}\delta \{p^2+q^4E\}\delta \{p\}𝒟p𝒟q^4}`$ $`={\displaystyle \delta \{q^4E\}𝒟q^4}`$ $`=1.`$ (16) Either of these choices imposes no restriction on $`E`$ whatsoever. Ignoring the nonphysical nature of the variables involved, one might possibly impose the condition $`{\displaystyle p𝑑q}=2\pi n,`$ (17) leading to a Bohr-Sommerfeld spectrum, which for this problem is incorrect. (The reader is encouraged to examine alternative choices of gauge.) Remark: It is instructive in this example to note that the Faddeev-Popov determinant $`\mathrm{\Delta }=\mathrm{\Pi }(4q^3)`$ and the reduced phase space is the single point $`(p,q)=(0,E^{1/4})`$. The point $`(p,q)=(0,E^{1/4})`$ corresponds to a Gribov copy. Clearly, in this case, reduction before quantization has led to the wrong result. Some workers may assert that such errors are merely “order $`\mathrm{}`$ corrections”. Although true, this argument cannot be used to defend the general procedure since the role of a quantization procedure, after all, should be to determine the correct spectrum for a specific problem, not a spectrum that is potentially incorrect even in its leading order. Examples of other work which arrive at the same conclusion are given in . ### 1.6 Outline of the remaining sections In the following section, Sec. 2, we present an overview of the projection operator approach to constrained system quantization with an emphasis on coherent-state representations. Section 3 deals with coherent-state path integrals without gauge fixing for closed first-class constrained systems. Extensions to general constraints such as open first-class or second-class systems are the subject of Sec. 4. Section 5 is devoted to selected examples of first-class systems, while Sec. 6 concentrates on two rather special applications. Finally, in Sec. 7 we comment on some other applications of the projection operator approach that have not been discussed in this paper. ## 2 OVERVIEW OF THE PROJECTION <br>OPERATOR APPROACH TO CON- <br>STRAINED SYSTEM QUANTIZATION ### 2.1 Coherent states Canonical quantization is consistent only for Cartesian phase space coordinates , and we assume that our original and unconstrained set of classical dynamical variables fulfill that condition. Then, for each classical coordinate $`q^j`$ and momentum $`p_j`$, $`1jJ`$, we may introduce associated self-adjoint canonical operators $`Q^j`$ and $`P_j`$, acting in a separable Hilbert space $``$, and which satisfy, in units where $`\mathrm{}=1`$, the canonical commutation relations $`[Q^j,P_k]=i\delta _k^j\mathrm{\hspace{0.17em}11}`$, with all other commutation relations vanishing. With the fiducial vector $`|0`$ a suitable normalized state—typically the ground state of a (unit-frequency) harmonic oscillator (but not always!)—we introduce the canonical coherent states (see, e.g., ) $`|p,qe^{iq^jP_j}e^{ip_jQ^j}|0,`$ (18) for all $`(p,q)\mathrm{IR}^{2J}`$, where $`p=\{p_j\}`$ and $`q=\{q^j\}`$. These states admit a resolution of unity in the form $`11={\displaystyle |p,qp,q|𝑑\mu (p,q)},d\mu (p,q)d^Jpd^Jq/(2\pi )^J,`$ (19) integrated over $`\mathrm{IR}^{2J}`$. The unit operator resides in the Hilbert space $``$ of the unconstrained system. We may conveniently represent this Hilbert space as follows. We first introduce the reproducing kernel $`p^{\prime \prime },q^{\prime \prime }|p^{},q^{}`$ as the overlap matrix element between any two coherent states. This expression is a bounded, continuous function that characterizes a (reproducing kernel Hilbert space) representation of $``$ appropriate to the unconstrained system as follows. A dense set of vectors in the associated functional Hilbert space is given by vectors of the form $`\psi (p,q)p,q|\psi ={\displaystyle \underset{l=1}{\overset{L}{}}}\alpha _lp,q|p_{(l)},q_{(l},`$ (20) for arbitrary sets $`\{\alpha _l\}`$ and $`\{p_{(l)},q_{(l)}\}`$ with $`L<\mathrm{}`$. The inner product of two such vectors is given by $`(\psi ,\xi ){\displaystyle \underset{l,m=1}{\overset{L,M}{}}}\alpha _l^{}\beta _mp_{(l)},q_{(l)}|\overline{p}_{(m)},\overline{q}_{(m)}`$ (21) $`={\displaystyle \psi (p,q)^{}\xi (p,q)𝑑\mu (p,q)},`$ (22) where $`\xi `$ is a second function defined in a manner analogous to $`\psi `$. A general vector in the functional Hilbert space is defined by a Cauchy sequence of such vectors, and all such vectors are given by bounded, continuous functions. The first form of the inner product applies in general only to vectors in the dense set, while the second form of the inner product holds for arbitrary vectors in the Hilbert space. We shall have more to say below regarding reproducing kernels and reproducing kernel Hilbert spaces. ### 2.2 Constraints Now suppose we introduce constraints into the quantum theory . In particular, we assume that $`\mathrm{EI}`$ denotes a projection operator onto the constraint subspace, i.e., the subspace on which the quantum constraints are satisfied (in a sense to be defined below), and which is called the physical Hilbert space $`_{phys}\mathrm{EI}`$. Later we shall discuss examples of $`\mathrm{EI}`$. Hence, if $`|\psi `$ denotes a general vector in the original (unconstrained) Hilbert space, the vector $`\mathrm{EI}|\psi _{phys}`$ represents its component within the physical subspace. As a Hilbert space, the physical subspace also admits a functional representation by means of a reproducing kernel which may be taken as $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$. In the same manner as before, it follows that a dense set of vectors in $`_{phys}`$ is given by functions of the form $`\psi (p,q)p,q|\mathrm{EI}|\psi ={\displaystyle \underset{l=1}{\overset{L}{}}}\alpha _lp,q|\mathrm{EI}|p_{(l)},q_{(l)},`$ (23) for arbitrary sets $`\{\alpha _l\}`$ and $`\{p_{(l)},q_{(l)}\}`$ with $`L<\mathrm{}`$. The inner product of two such vectors is given by $`(\psi ,\xi ){\displaystyle \underset{l,m=1}{\overset{L,M}{}}}\alpha _l^{}\beta _mp_{(l)},q_{(l)}|\mathrm{EI}|\overline{p}_{(m)},\overline{q}_{(m)}`$ $`={\displaystyle \psi (p,q)^{}\xi (p,q)𝑑\mu (p,q)}.`$ (24) Again, a general vector in the functional Hilbert space is defined by means of a Cauchy sequence, and all such vectors are given by bounded, continuous functions. Note well, in the case illustrated, that even though $`\mathrm{EI}`$, the functional representation of the unconstrained and the constrained Hilbert spaces are identical, namely by functions of $`(p,q)\mathrm{IR}^{2J}`$, and the form of the inner product is identical in the two cases. This situation holds even if $`_{phys}`$ is one dimensional! The relation between the self-adjoint constraint operators $`\mathrm{\Phi }_a`$, $`1aA`$, $`A<\mathrm{}`$, and the projection operator $`\mathrm{EI}`$ may take several different forms. Unless otherwise specified, we shall assume that $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2`$ is self adjoint and that $`\mathrm{EI}=\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2\delta (\mathrm{})^2),`$ (25) where $`\delta =\delta (\mathrm{})`$ (not a Dirac $`\delta `$-function!) is a regularization parameter which is chosen in accord with rules to be discussed below. ### 2.3 Dynamics for first-class systems Suppose further that the Hamiltonian $``$ respects the first-class character of the constraints. It follows in this case that $`[\mathrm{EI},]=0`$ or stated otherwise that $`e^{it}\mathrm{EI}\mathrm{EI}e^{it}\mathrm{EI}\mathrm{EI}e^{i(\mathrm{EI}\mathrm{EI})t}\mathrm{EI}.`$ (26) Dynamics in the physical subspace is then fully determined by the propagator on $`_{phys}`$, which is given in the relevant functional representation by $`p^{\prime \prime },q^{\prime \prime }|e^{it}\mathrm{EI}|p^{},q^{}.`$ (27) In (27) we have achieved a fully gauge invariant propagator without having to reduce the range or even the number of the original classical variables nor change the original form of the inner product on the functional Hilbert space representation. Any observable $`𝒪`$$``$ included—satifies $`[\mathrm{EI},𝒪]=0`$, and relations similar to (26) follow with $``$ replaced by $`𝒪`$. ### 2.4 Zero in the continuous spectrum The foregoing scenario has assumed that the appropriate $`_{phys}`$ is given by means of a projection operator $`\mathrm{EI}`$ acting on the original Hilbert space. This situation holds true whenever the set of quantum constraints admits zero as a common point in their discrete spectrum; in that case $`\mathrm{EI}`$ defines the subspace where the constraints all vanish. That situation may not always hold true, but even in case zero lies in the continuous spectrum for some or all of the constraints, a suitable result may generally be given by matrix elements of a sequence of rescaled projection operators, say $`c_\delta \mathrm{EI}`$, $`c_\delta >0`$, as $`\delta 0`$. Specifically, we consider the limit of a sequence of reproducing kernels $`c_\delta p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$, which—if the limit is a nonvanishing continuous function—defines a new reproducing kernel, and thereby a new reproducing kernel Hilbert space, within which the appropriate constraints are fulfilled. In such a limit certain variables may cease to be relevant and as a consequence the local integral representation of the inner product, if any, may require modification. On the other hand, the definition of the inner product by sums involving the reproducing kernel will always hold. We refer to the result of such a limiting operation as a reduction of the reproducing kernel. A simple example should help clarify what we mean by a reduction of the reproducing kernel. Consider the example $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$ $`=\pi ^{1/2}{\displaystyle _\delta ^\delta }\mathrm{exp}[\frac{1}{2}(kp^{\prime \prime })^2+ik(q^{\prime \prime }q^{})\frac{1}{2}(kp^{})^2]dk,`$ (28) where $`\mathrm{EI}=\mathrm{EI}(P^2\delta ^2)`$, which defines a reproducing kernel for any $`\delta >0`$ that corresponds to an infinite dimensional Hilbert space. (If $`\delta =\mathrm{}`$ the result is the usual canonical coherent state overlap and characterizes the unconstrained Hilbert space.) If we take the limit of the expression as it stands as $`\delta 0`$, the result will vanish. What we need to do is extract the germ of the projection operator as we let $`\delta `$ go to zero. Therefore, let us first multiply this expression by $`\pi ^{1/2}/(2\delta )`$ \[$`c_\delta `$ in this case\] and take the limit $`\delta 0`$. The result is the expression $`𝒦(p^{\prime \prime };p^{})=e^{{\scriptscriptstyle \frac{1}{2}}\left(p^{\prime \prime 2}+p^2\right)},`$ (29) which has become a reproducing kernel that characterizes a one-dimensional Hilbert space with every functional representative proportional to $`\chi _o(p)\mathrm{exp}(p^2/2)`$. This one-dimensional Hilbert space representation also admits a local integral representation for the inner product given by $`(\chi ,\chi )=|\chi (p)|^2𝑑p/\sqrt{\pi }.`$ (30) In the present case, it is clear that one may reduce the reproducing kernel even further by choosing $`p=c`$, an arbitrary but fixed constant. This kind of reduction—in which the latter reproducing kernel Hilbert space is equivalent to the former reproducing kernel Hilbert space—is analogous to choosing a gauge in the classical theory. We shall see another example of this latter kind of reduction later. The example presently under discussion is also an important one inasmuch as it illustrates how a constraint operator with its zero lying in the continuous spectrum is dealt with in the coherent-state, projection-operator approach. Some other approaches to deal with the problem of zero in the continuous spectrum may be traced from . ### 2.5 Alternative view of continuous zeros If $`\delta 1`$ in (28), then it may be approximately evaluated as $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$ $`=\pi ^{1/2}\delta e^{{\scriptscriptstyle \frac{1}{2}}\left(p^{\prime \prime 2}+p^2\right)}{\displaystyle \frac{\mathrm{sin}[\delta (q^{\prime \prime }q^{})]}{\delta (q^{\prime \prime }q^{})}}+O(\delta ^2).`$ (31) When $`\delta =10^{1000}`$, or some other extremely tiny factor, it is clear that for all practical purposes it is sufficient to accept just the first term in (31), ignoring the term $`O(\delta ^2)`$, as the “reduced” reproducing kernel. The resultant expression is indeed a proper reproducing kernel for which inner products are given with the full set of integration variables and the normal integration range. So long as $`q`$ values are “normal sized”, e.g., $`|q|<10^{500}`$ in the present case, there is no practical distinction between the space of functions generated by (29) and that generated by (31). In other words, if $`\delta `$ is chosen extremely close to zero, but still positive, it is not actually necessary to take the limit $`\delta 0`$ in order to do practical calculations. Even though this is the case, we shall for the most part in the examples we study take a full reduction by first rescaling the reproducing kernel (by an appropriate factor $`c_\delta `$) and then taking the limit $`\delta 0`$. ## 3 COHERENT STATE PATH INTEGRALS <br>WITHOUT GAUGE FIXING As introduced above, canonical coherent states may be defined by the relation $`|p,qe^{iq^jP_j}e^{ip_jQ^j}|0,`$ (32) for all $`(p,q)`$, where the fiducial vector $`|0`$ traditionally denotes a normalized, unit frequency, harmonic oscillator ground state, and the coherent states admit a resolution of unity in the form $`11=|p,qp,q|𝑑\mu (p,q),d\mu (p,q)d^Jpd^Jq/(2\pi )^J,`$ (33) where the integration is over $`\mathrm{IR}^{2J}`$. Note that the integration domain and the form of the measure are unique. Based on such coherent states, we introduce the upper symbol for a general operator $`(P,Q)`$, $`H(p,q)p,q|(P,Q)|p,q=p,q|:H(P,Q):|p,q`$ (34) which is related to the normal-ordered form as shown. (N.B. Some workers would call $`H(p,q)`$ the lower symbol.) If $`(P,Q)`$ denotes the quantum Hamiltonian, then we shall adopt $`H(p,q)`$ as the classical Hamiltonian. We also note that an important one-form generated by the coherent states is given by $`ip,q|d|p,q=p_jdq^j`$. Using these quantities, and the time ordering operator $`𝖳`$, the coherent state path integral for the propagator generated by the time-dependent Hamiltonian $`(P,Q)+\lambda ^a(t)\mathrm{\Phi }_a(P,Q)`$ is readily given by $`p^{\prime \prime },q^{\prime \prime }|𝖳e^{i_0^T[(P,Q)+\lambda ^a(t)\mathrm{\Phi }_a(P,Q)]𝑑t}|p^{},q^{}`$ $`=\underset{ϵ0}{lim}{\displaystyle \underset{l=0}{\overset{N}{}}p_{l+1},q_{l+1}|e^{iϵ(+\lambda _l^a\mathrm{\Phi }_a)}|p_l,q_l\underset{l=1}{\overset{N}{}}d\mu (p_l,q_l)}`$ $`={\displaystyle \mathrm{exp}\{i[ip,q|(d/dt)|p,qp,q|+\lambda ^a(t)\mathrm{\Phi }_a|p,q]𝑑t\}𝒟\mu (p,q)}`$ $`={\displaystyle \mathrm{exp}\{i[p_j\dot{q}^jH(p,q)\lambda ^a(t)\varphi _a(p,q)]𝑑t\}𝒟p𝒟q}.`$ (35) Here, in the second line, we have set $`ϵT/(N+1)`$, made a Trotter-product like approximation to the evolution operator, repeatedly inserted the resolution of unity, and set $`p_{N+1},q_{N+1}=p^{\prime \prime },q^{\prime \prime }`$ and $`p_0,q_0=p^{},q^{}`$. In the third and fourth lines we have formally interchanged the continuum limit and the integrations, and written for the integrand the form it would assume for continuous and differentiable paths ($``$ denotes a formal normalization constant). The result evidently depends on the chosen form of the functions $`\{\lambda ^a(t)\}`$. ### 3.1 Enforcing the quantum constraints Let us next introduce the quantum analog of the initial value equation. For simplicity we assume that the constraint operators form a compact group; more general situations are dealt with below. In that case $`\mathrm{EI}e^{i\xi ^a\mathrm{\Phi }_a(P,Q)}\delta \xi =\mathrm{EI}(\mathrm{\Phi }_a=0,\mathrm{\hspace{0.33em}\hspace{0.33em}1}aA)=\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2=0)`$ (36) defines a projection operator onto the subspace for which $`\mathrm{\Phi }_a=0`$ provided that $`\delta \xi `$ denotes the normalized, $`\delta \xi =1`$, group invariant measure. It follows from (36) that $`e^{i\tau ^a\mathrm{\Phi }_a}\mathrm{EI}=\mathrm{EI}.`$ (37) We now project the propagator (3) onto the quantum constraint subspace which leads to the following set of relations $`{\displaystyle p^{\prime \prime },q^{\prime \prime }|𝖳e^{i{\scriptscriptstyle [+\lambda ^a(t)\mathrm{\Phi }_a]𝑑t}}|\overline{p}^{},\overline{q}^{}\overline{p}^{},\overline{q}^{}|\mathrm{EI}|p^{},q^{}𝑑\mu (\overline{p}^{},\overline{q}^{})}`$ $`=p^{\prime \prime },q^{\prime \prime }|𝖳e^{i{\scriptscriptstyle [+\lambda ^a(t)\mathrm{\Phi }_a]𝑑t}}\mathrm{EI}|p^{},q^{}`$ $`=limp^{\prime \prime },q^{\prime \prime }|[{\displaystyle \underset{l}{\overset{}{}}}(e^{iϵ}e^{iϵ\lambda _l^a\mathrm{\Phi }_a})]\mathrm{EI}|p^{},q^{}`$ $`=p^{\prime \prime },q^{\prime \prime }|e^{iT}e^{i\tau ^a\mathrm{\Phi }_a}\mathrm{EI}|p^{},q^{}`$ $`=p^{\prime \prime },q^{\prime \prime }|e^{iT}\mathrm{EI}|p^{},q^{},`$ (38) where $`\tau ^a`$ incorporates the functions $`\lambda ^a`$ as well as the structure parameters $`c_{ab}^c`$ and $`h_a^b`$. Alternatively, this expression has the formal path integral representation $`{\displaystyle \mathrm{exp}\{i[p_j\dot{q}^jH(p,q)\lambda ^a(t)\varphi _a(p,q)]𝑑ti\xi ^a\varphi _a(p^{},q^{})\}𝒟\mu (p,q)\delta \xi }.`$ (39) On comparing (3) and (39), we observe that after projection onto the quantum constraint subspace the propagator is entirely independent of the choice of the Lagrange multiplier functions. In other words, the projected propagator is gauge invariant. We may also express the physical (projected) propagator in a more general form, namely, $`{\displaystyle \mathrm{exp}\{i[p_j\dot{q}^jH(p,q)\lambda ^a(t)\varphi _a(p,q)]𝑑t\}𝒟\mu (p,q)𝒟C(\lambda )}`$ $`=p^{\prime \prime },q^{\prime \prime }|e^{iT}\mathrm{EI}|p^{},q^{}`$ (40) provided that $`𝒟C(\lambda )=1`$ and that such an average over the functions $`\{\lambda ^a(t)\}`$ introduces (at least) one factor $`\mathrm{EI}`$. ### 3.2 Reproducing kernel Hilbert spaces The coherent-state matrix elements of $`\mathrm{EI}`$ define a fundamental kernel $`𝒦(p^{\prime \prime },q^{\prime \prime };p^{},q^{})p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{},`$ (41) which is a bounded, continuous function for any projection operator $`\mathrm{EI}`$, especially including the unit operator. It follows that $`𝒦(p^{\prime \prime },q^{\prime \prime };p^{},q^{})^{}=𝒦(p^{},q^{};p^{\prime \prime },q^{\prime \prime })`$ as well as $`{\displaystyle \underset{k,l=1}{\overset{K}{}}}\alpha _k^{}\alpha _l𝒦(p_k,q_k;p_l,q_l)0`$ (42) for all sets $`\{\alpha _k\}`$, $`\{(p_k,q_k)\}`$, and all $`K<\mathrm{}`$. The last relation is an automatic consequence of the complex conjugate property and the fact that $`𝒦(p^{\prime \prime },q^{\prime \prime };p^{},q^{})={\displaystyle 𝒦(p^{\prime \prime },q^{\prime \prime };p,q)𝒦(p,q;p^{},q^{})𝑑\mu (p,q)}`$ (43) holds in virtue of the coherent state resolution of unity and the properties of $`\mathrm{EI}`$. As noted earlier, the function $`𝒦`$ is called the reproducing kernel and the Hilbert space it engenders is termed a reproducing kernel Hilbert space . A dense set of elements in the Hilbert space is given by functions of the form $`\psi (p,q)={\displaystyle \underset{k=1}{\overset{K}{}}}\alpha _k𝒦(p,q;p_k,q_k),`$ (44) and the inner product of this function has two equivalent forms given by $`(\psi ,\psi )={\displaystyle \underset{k,l=1}{\overset{K}{}}}\alpha _k^{}\alpha _l𝒦(p_k,q_k;p_l,q_l)`$ (45) $`=\psi (p,q)^{}\psi (p,q)𝑑\mu (p,q).`$ (46) The inner product of two distinct functions may be determined by polarization of the norm squared . Clearly, the entire Hilbert space is characterized by the reproducing kernel $`𝒦`$. Change the kernel $`𝒦`$ and one changes the representation of the Hilbert space. Following a suitable limit of the kernel $`𝒦`$, it is even possible to change the dimension of the Hilbert space, as already illustrated earlier. ### 3.3 Reduction of the reproducing kernel Suppose the reproducing kernel depends on a number of variables and additional parameters. We can generate new reproducing kernels from a given kernel by a variety of means. For example, the expressions $`𝒦_1(p^{\prime \prime };p^{})=𝒦(p^{\prime \prime },c;p^{},c),`$ (47) $`𝒦_2(p^{\prime \prime };p^{})=f(q^{\prime \prime })^{}f(q^{})𝒦(p^{\prime \prime },q^{\prime \prime };p^{},q^{})𝑑q^{\prime \prime }𝑑q^{},`$ (48) $`𝒦_3(p^{\prime \prime },q^{\prime \prime };p^{},q^{})=lim𝒦(p^{\prime \prime },q^{\prime \prime };p^{},q^{})`$ (49) each generate a new reproducing kernel provided the resultant function remains continuous. In general, however, the inner product in the Hilbert space generated by the new reproducing kernel is only given by an analog of (21) and not by (22), although frequently some sort of local integral representation for the inner product may exist. Let us offer an example of the reduction of a reproducing kernel that is a slight generalization of the earlier example. Let the expression $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$ $`\pi ^{J/2}{\displaystyle _\delta ^\delta }\mathrm{}{\displaystyle _\delta ^\delta }\mathrm{exp}[\frac{1}{2}(kp^{\prime \prime })^2+ik(q^{\prime \prime }q^{})\frac{1}{2}(kp^{})^2]d^Jk`$ (50) denote a reproducing kernel for any $`\delta >0`$. In the present case it follows that $`\mathrm{EI}\mathrm{\Pi }_{j=1}^J\mathrm{EI}(\delta P_j\delta )`$. When $`\delta 0`$, then (3.3) vanishes. However, if we first multiply by $`\delta ^J`$—or more conveniently by $`\pi ^{J/2}(2\delta )^J`$—before taking the limit, the result becomes $`\underset{\delta 0}{lim}\pi ^{J/2}(2\delta )^Jp^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}=\mathrm{exp}(\frac{1}{2}p^{\prime \prime \mathrm{\hspace{0.17em}2}})\mathrm{exp}(\frac{1}{2}p^{\mathrm{\hspace{0.17em}2}}),`$ (51) which is continuous and therefore denotes the reproducing kernel for some Hilbert space. Note that the classical variables $`q^{\prime \prime }`$ and $`q^{}`$ have disappeared, which on reference to (32) implies that all “$`P_j=0`$”. In the present example, the resultant Hilbert space is one dimensional, and the inner product may be given either by a sum as in (21) involving the $`p`$ variables alone or by a local integral representation now using the measure $`\pi ^{J/2}d^Jp`$, namely, $`(\chi ,\chi )=|\chi (p)|^2\pi ^{J/2}d^Jp.`$ (52) This example illustrates the case where the constraints are “$`P_j=0`$”, for all $`j`$, a situation where zero lies in the continuous spectrum. We may also use this example to illustrate how several constraints may be replaced by a single constraint. The several constraints “$`P_j=0`$”, for all $`j`$, were first approximated by the regularized constraints $`P_j^2\delta ^2`$, $`\delta >0`$, for all $`j`$. Alternatively, we may also regularize the constraints in the form $`\mathrm{\Sigma }_jP_j^2\delta ^2`$. Furthermore, if we use $`\mathrm{EI}=\mathrm{EI}(\mathrm{\Sigma }_jP_j^2\delta ^2)`$, then it is clear that a new prefactor, also proportional to $`\delta ^J`$, can be chosen so that (51) again emerges as $`\delta 0`$. ### 3.4 Single regularized constraints Clearly, the set of real classical constraints $`\varphi _a=0`$, $`1aA`$, is equivalent to the single classical constraint $`\mathrm{\Sigma }_a\varphi _a^2=0`$. Likewise, the set of (idealized) quantum constraints “$`\mathrm{\Phi }_a|\psi _{phys}=0`$”, $`1aA`$, where each $`\mathrm{\Phi }_a`$ is self adjoint, is equivalent to the single (idealized) quantum constraint “$`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2|\psi _{phys}=0`$”, where we further assume that $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2`$ is a self-adjoint operator. In general, however, the only solution of the idealized quantum constraint is the zero vector, $`|\psi _{phys}=0`$. To overcome this difficulty, we relax the idealized quantum constraint and instead generally adopt the regularized form of the constraint given by $`|\psi _{phys}_{phys}\mathrm{EI}`$, where $`\mathrm{EI}=\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2\delta (\mathrm{})^2).`$ (53) Here $`\delta (\mathrm{})`$ is a regularization parameter and the inequality means that in a spectral resolution of $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2_0^{\mathrm{}}\lambda 𝑑E(\lambda )`$ that $`\mathrm{EI}{\displaystyle _0^{\delta (\mathrm{})^2}}𝑑E(\lambda )=E(\delta (\mathrm{})^2).`$ (54) Let us examine three basic examples. First, let zero be in the discrete spectrum of $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2`$. Then, it follows that there exists a $`\delta _1(\mathrm{})^2`$ such that for all $`\delta (\mathrm{})^2`$, $`0<\delta (\mathrm{})^2<\delta _1(\mathrm{})^2`$, then $`\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2\delta (\mathrm{})^2)=\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2=0)`$. Second, if $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2`$ has its zero in the continuum, then $`\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2\delta ^2)`$ is infinite dimensional for all $`\delta >0`$, but $`\mathrm{EI}`$ vanishes weakly as $`\delta 0`$. For such cases we consider $`c_\delta \mathrm{EI}`$ and choose the sequence $`c_\delta `$ to weakly extract the germ of $`\mathrm{EI}`$ as $`\delta 0`$, just as in the examples illustrated above. Third, in a case to be studied later, suppose that zero is not in the spectrum of the operator $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2`$. Since $`\mathrm{\Sigma }_a\varphi _a^2=0`$ classically, it follows that spectral values of $`\mathrm{\Sigma }_a\mathrm{\Phi }_a^2`$ are $`o(\mathrm{}^0)`$ close to zero. A relevant example discussed later is where $`\mathrm{\Phi }_1=P`$ and $`\mathrm{\Phi }_2=Q`$. Then $`\mathrm{EI}(P^2+Q^2\mathrm{})=|00|`$ is a one-dimensional projection operator onto the harmonic oscillator ground state $`|0`$. Observe in this case that $`\delta (\mathrm{})^2=\mathrm{}`$, which vanishes when $`\mathrm{}0`$; note also that we cannot reduce this parameter further since $`\mathrm{EI}(P^2+Q^2<\mathrm{})0`$. Thus, in some cases, whether we use “$``$” or “$`<`$” in the inequality defining the projection operator can make a real difference. The three types of examples discussed above illustrate three qualitatively different behaviors possible for the projection operator $`\mathrm{EI}`$. As we proceed, we shall find the use of a single regularized constraint will be an important unifying principle in treating the most general multiple constraint situation imaginable. ### 3.5 Basic first-class constraint example Consider the system with two degrees of freedom, a vanishing Hamiltonian, and a single constraint, characterized by the action $`I=[\frac{1}{2}(p_1\dot{q}_1q_1\dot{p}_1+p_2\dot{q}_2q_2\dot{p}_2)\lambda (q_2p_1p_2q_1)]dt,`$ (55) where for notational convenience we have lowered the index on the $`q`$ variables. Note that we have chosen a different form for the kinematic part of the action which amounts to a change of phase for the coherent states, and in particular a factor of $`e^{ipq/2}`$ has been introduced on the right side of (18), or, equivalently, both generators appear in the same exponent. It follows that $`{\displaystyle \mathrm{exp}\{i[\frac{1}{2}(p_1\dot{q}_1q_1\dot{p}_1+p_2\dot{q}_2q_2\dot{p}_2)\lambda (q_2p_1p_2q_1)]𝑑t\}}`$ $`\times 𝒟p𝒟q𝒟C(\lambda )`$ $`=p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{},`$ (56) where we choose $`\mathrm{EI}=(2\pi )^1{\displaystyle _0^{2\pi }}e^{i\xi (Q_2P_1P_2Q_1)}𝑑\xi =\mathrm{EI}(L_3=0).`$ (57) Based on the fact that $`p^{\prime \prime },q^{\prime \prime }|p^{},q^{}=\mathrm{exp}(\frac{1}{2}|z_1^{\prime \prime }|^2\frac{1}{2}|z_2^{\prime \prime }|^2+z_1^{\prime \prime }z_1^{}+z_2^{\prime \prime }z_2^{}\frac{1}{2}|z_1^{}|^2\frac{1}{2}|z_2^{}|^2),`$ (58) where $`z_1^{}(q_1^{}+ip_1^{})/\sqrt{2}`$, etc., it is straightforward to show that $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}=\mathrm{exp}(\frac{1}{2}|z_1^{\prime \prime }|^2\frac{1}{2}|z_2^{\prime \prime }|^2\frac{1}{2}|z_1^{}|^2\frac{1}{2}|z_2^{}|^2)`$ $`\times I_0((z_1^{\prime \prime 2}+z_2^{\prime \prime 2})^{1/2}(z_1^2+z_2^2)^{1/2}),`$ (59) with $`I_0`$ a standard Bessel function. We emphasize again that although the Hilbert space has been strictly reduced by the introduction of $`\mathrm{EI}`$, the reproducing kernel (3.5) leads to a reproducing kernel Hilbert space with an inner product having the same number of integration variables and domain of integration as in the unconstrained case. ## 4 APPLICATION TO GENERAL <br>CONSTRAINTS ### 4.1 Classical considerations When dealing with a general constraint situation it will typically happen that the self-consistency of the equations of motion may determine some or all of the Lagrange multipliers in order for the system to remain on the classical constraint hypersurface. For example, if the Hamiltonian attempts to force points initially lying on the constraint hypersurface to leave that hypersurface, then the Lagrange multipliers must supply the necessary forces for the system to remain on the constraint hypersurface. We may elaborate on this situation as follows. Since $`\varphi _a(p,q)=0`$ for all $`a`$ defines the constraint hypersurface, it is also necessary, for all $`a`$, that $`\dot{\varphi }_a(p,q)\{\varphi _a(p,q),H(p,q)\}+\lambda ^b(t)\{\varphi _a(p,q),\varphi _b(p,q)\}0`$ (60) also holds on the constraint hypersurface. If the Poisson brackets fulfill the conditions given in (3) and (4), then it follows that $`\dot{\varphi }_a(p,q)0`$ on the constraint hypersurface for any choice of the Lagrange multipliers $`\{\lambda ^a(t)\}`$. This is the case for first-class constraints, and to obtain specific solutions to the dynamical equations it is necessary to specify some choice of the Lagrange multipliers, i.e., to select a gauge. However, if (3), or (3) and (4) do not hold on the constraint hypersurface, the situation changes. For example, let us first assume that (4) holds but that $`\mathrm{\Delta }_{ab}(p,q)\{\varphi _a(p,q),\varphi _b(p,q)\}`$ (61) is a nonsingular matrix on the constraint hypersurface. In this case it follows that we must choose $`\lambda ^a(t)0`$ for all $`a`$ to satisfy (60). More generally, we must choose $`\lambda ^a(t)(\mathrm{\Delta }^1(p,q))^{ab}\{\varphi _b(p,q),H(p,q)\}`$ (62) in order that (60) will be satisfied. When the Lagrange multipliers are not arbitrary but rather must be specifically chosen in order to keep the system on the constraint hypersurface, then we say that we deal with second-class constraints. Of course, there are also intermediate situations where part of the constraints are first class while some are second class; in this case the matrix $`\mathrm{\Delta }_{ab}(p,q)`$ would be singular but would have a nonzero rank on the constraint hypersurface. Remark: It is useful to also imagine solving the differential equation (60) as a computer might do it, namely, by an iteration procedure. In particular, we could imagine evolving by a small time step $`ϵ`$ by the first (Hamiltonian) term, then using the second (constraint) term to choose $`\lambda ^a`$ at that moment to force the system back onto the constraint hypersurface, and afterwards continuing this procedure over and over. A proper solution can be obtained this way by taking the limit of these approximate solutions as $`ϵ0`$. An analogue of this procedure will be used in our quantum discussion. There is also a third situation that may arise, namely constraints that are first class from a classical point of view but are second class quantum mechanically. Such constraints would arise if $`\mathrm{\Delta }_{ab}(p,q)=Y_{ab}^c(p,q)\varphi _c(p,q),`$ (63) where, for the sake of convenience, we assume that the quantities $`Y_{ab}^c(p,q)`$ are all uniformly bounded away from zero and infinity, i.e., $`0<CY_{ab}^c(p,q)`$ $`D<\mathrm{}`$. In that case $`\mathrm{\Delta }_{ab}(p,q)`$ would vanish on the constraint hypersurface classically. Quantum mechanically, the expression for the commutator is proportional to $`\mathrm{}`$ and may be taken as $`i[\mathrm{\Phi }_a(P,Q),\mathrm{\Phi }_b(P,Q)]=\frac{1}{2}[Y_{ab}^c(P,Q)\mathrm{\Phi }_c(P,Q)+\mathrm{\Phi }_c(P,Q)Y_{ab}^c(P,Q)].`$ (64) If we assume that “$`\mathrm{\Phi }_a(P,Q)|\psi _{phys}=0`$”, then self-consistency requires that “$`[\mathrm{\Phi }_c(P,Q),Y_{ab}^c(P,Q)]|\psi _{phys}=0`$”, an expression which is now proportional to $`\mathrm{}^2`$. If this expression vanishes it causes no problem; if it does not vanish one says that there is a “factor ordering problem” or an “anomaly”. As Jackiw has often stressed, it would be preferable to call an anomaly “quantum mechanical symmetry breaking”, a phrase which more accurately describes what it is and what it does. Whatever it is called, the resultant quantum constraints are second class even though they were classically first class. As is well known, gravity falls into just this category. In this section we take up the quantization of these more general situations involving both first and second class constraints . ### 4.2 Quantum considerations As in previous sections, we let $`\mathrm{EI}`$ denote the projection operator onto the quantum constraint subspace. Motivated by the classical comments given above we consider the quantity $`limp^{\prime \prime },q^{\prime \prime }|\mathrm{EI}e^{iϵ}\mathrm{EI}e^{iϵ}\mathrm{}\mathrm{EI}e^{iϵ}\mathrm{EI}|p^{},q^{}`$ (65) where the limit, as usual, is for $`ϵ0`$. The physics behind this expression is as follows. Reading from right to left we first impose the quantum initial value equation, and then propagate for a small amount of time ($`ϵ`$). Next we recognize that the system may have left the quantum constraint subspace, and so we project it back onto that subspace, and so on over and over. In the limit that $`ϵ0`$ the system remains within the quantum constraint subspace and (65) actually leads to $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}e^{iT(\mathrm{EI}\mathrm{EI})}\mathrm{EI}|p^{},q^{},`$ (66) which clearly illustrates temporal evolution entirely within the quantum constraint subspace. If we assume that $`\mathrm{EI}\mathrm{EI}`$ is a self-adjoint operator, then we conclude that (66) describes a unitary time evolution within the quantum constraint subspace. The expression (65) may be developed in two additional and alternative ways. First, we repeatedly insert the resolution of unity in such a way that (65) becomes $`lim{\displaystyle \underset{l=0}{\overset{N}{}}p_{l+1},q_{l+1}|\mathrm{EI}e^{iϵ}\mathrm{EI}|p_l,q_l\underset{l=1}{\overset{N}{}}d\mu (p_l,q_l)}.`$ (67) We wish to turn this expression into a formal path integral, but the procedure used previously relied on the use of unit vectors, and the vectors $`\mathrm{EI}|p,q`$ are generally not unit vectors. Thus, let us rescale the factors in the integrand introducing $`|p,q\mathrm{EI}|p,q/\mathrm{EI}|p,q`$ (68) which are unit vectors. If we let $`M^{\prime \prime }\mathrm{EI}|p^{\prime \prime },q^{\prime \prime }`$, $`M^{}\mathrm{EI}|p^{},q^{}`$, and observe that $`\mathrm{EI}|p,q^2=p,q|\mathrm{EI}|p,q`$, it follows that (67) may be rewritten as $`M^{\prime \prime }M^{}lim{\displaystyle \underset{l=0}{\overset{N}{}}p_{l+1},q_{l+1}|e^{iϵ}|p_l,q_l\underset{l=1}{\overset{N}{}}p_l,q_l|\mathrm{EI}|p_l,q_ld\mu (p_l,q_l)}.`$ (69) This expression is represented by the formal path integral $`M^{\prime \prime }M^{}{\displaystyle \mathrm{exp}\{i[ip,q|(d/dt)|p,qp,q||p,q]𝑑t\}𝒟_E\mu (p,q)},`$ (70) where the new formal measure for the path integral is defined in an evident fashion from its lattice prescription. We can also reexpress this formal path integral in terms of the original bra and ket vectors in the form $`M^{\prime \prime }M^{}{\displaystyle }\mathrm{exp}\{i[ip,q|\mathrm{EI}(d/dt)\mathrm{EI}|p,q/p,q|\mathrm{EI}|p,q`$ $`p,q|\mathrm{EI}\mathrm{EI}|p,q/p,q|\mathrm{EI}|p,q]dt\}𝒟_E\mu (p,q).`$ (71) This last relation concludes our second route of calculation beginning with (65). The third relation we wish to derive uses an integral representation for the projection operator $`\mathrm{EI}`$ generally given by $`\mathrm{EI}=e^{i\xi ^a\mathrm{\Phi }_a(P,Q)}f(\xi )\delta \xi `$ (72) for a suitable function $`f`$. Thus we rewrite (65) in the form $`lim{\displaystyle p^{\prime \prime },q^{\prime \prime }|e^{iϵ\lambda _N^a\mathrm{\Phi }_a}e^{iϵ}e^{iϵ\lambda _{N1}^a\mathrm{\Phi }_a}e^{iϵ}\mathrm{}e^{iϵ\lambda _1^a\mathrm{\Phi }_a}e^{iϵ}e^{iϵ\lambda _0^a\mathrm{\Phi }_a}|p^{},q^{}}`$ $`\times f(ϵ\lambda _N)\mathrm{}f(ϵ\lambda _0)\delta ϵ\lambda _N\mathrm{}\delta ϵ\lambda _0.`$ (73) Next we insert the coherent-state resolution of unity at appropriate places to find that (4.2) may also be given by $`lim{\displaystyle p_{N+1},q_{N+1}|e^{iϵ\lambda _N^a\mathrm{\Phi }_a}|p_N,q_N\underset{l=0}{\overset{N1}{}}p_{l+1},q_{l+1}|e^{iϵ}e^{iϵ\lambda _l^a\mathrm{\Phi }_a}|p_l,q_l}`$ $`\times [{\displaystyle \underset{l=1}{\overset{N}{}}}d\mu (p_l,q_l)f(ϵ\lambda _l)\delta ϵ\lambda _l]f(ϵ\lambda _0)\delta ϵ\lambda _0.`$ (74) Following the normal pattern, this last expression may readily be turned into a formal coherent-state path integral given by $`{\displaystyle \mathrm{exp}\{i[p_j\dot{q}^jH(p,q)\lambda ^a(t)\varphi _a(p,q)]𝑑t\}𝒟\mu (p,q)𝒟E(\lambda )},`$ (75) where $`E(\lambda )`$ is a measure designed so as to insert the projection operator $`\mathrm{EI}`$ at every time slice. This usage of the Lagrange multipliers to ensure that the quantum system remains within the quantum constraint subspace is similar to their usage in the classical theory to ensure that the system remains on the classical constraint hypersurface. On the other hand, it is also possible to use the measure $`E(\lambda )`$ in the case of closed first-class constraints as well; this would be just one of the acceptable choices for the measure $`C(\lambda )`$ designed to put at least one projection operator $`\mathrm{EI}`$ into the propagator. In summary, we have established the equality of the three expressions $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}e^{iT(\mathrm{EI}\mathrm{EI})}\mathrm{EI}|p^{},q^{}`$ $`=M^{\prime \prime }M^{}{\displaystyle }\mathrm{exp}\{i[ip,q|\mathrm{EI}(d/dt)\mathrm{EI}|p,q/p,q|\mathrm{EI}|p,q`$ $`p,q|\mathrm{EI}\mathrm{EI}|p,q/p,q|\mathrm{EI}|p,q]dt\}𝒟_E\mu (p,q)`$ $`={\displaystyle \mathrm{exp}\{i[p_j\dot{q}^jH(p,q)\lambda ^a(t)\varphi _a(p,q)]𝑑t\}𝒟\mu (p,q)𝒟E(\lambda )}.`$ (76) This concludes our initial derivation of path integral formulas for general constraints. Observe that we have not introduced any $`\delta `$-functionals, nor, in the middle expression, reduced the number of integration variables or the limits of integration in any way even though in that expression the integral over the Lagrange multipliers has been carried out. ### 4.3 Universal procedure to generate single regularized constraints The preceding section developed a functional integral approach suitable for a general set of constraints, but it had one weak point, namely, it required prior knowledge of the constraints themselves in order to choose $`f(\xi )`$ in (72) so as to construct the appropriate projection operator. Is there any way to construct $`\mathrm{EI}`$ without prior knowledge of the form the constraints will take? The answer is yes! We first observe that the evolution operator appearing in (3) may be written in the form of a lattice limit given by $`\underset{ϵ0}{lim}{\displaystyle \underset{1nN}{\overset{}{}}}\left[𝖳e^{i_{(n1)ϵ}^{nϵ}(t)𝑑t}\right]\left[𝖳e^{i_{(n1)ϵ}^{nϵ}\lambda ^a(t)\mathrm{\Phi }_a𝑑t}\right],`$ (77) where $`ϵT/N`$ and the directed product (symbol $``$) also respects the time ordering. Thus, this expression is simply an alternating sequence of short-time evolutions, first by $`\lambda ^a(t)\mathrm{\Phi }_a`$, second by $`(t)`$, a pattern which is then repeated $`N1`$ more times. The validity of this Trotter-product form follows whenever $`(t)^2+\mathrm{\Phi }_a\delta ^{ab}\mathrm{\Phi }_b`$ is essentially self adjoint for all $`t`$, $`0tT`$. As a slight generalization, we shall assume that $`(t)^2+\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_b`$ is essentially self adjoint for all $`t`$, $`0tT`$. Here the real, symmetric coefficients $`M^{ab}(=M^{ba})`$ are the elements of a positive-definite matrix, i.e., $`\{M^{ab}\}>0`$. For a finite number of constraints, $`A<\mathrm{}`$, it is sufficient to assume that $`M^{ab}=\delta ^{ab}`$. Other choices for $`M^{ab}`$ may be relevant when $`A=\mathrm{}`$. (We do not explicitly consider the case $`A=\mathrm{}`$ in this article; for some examples see .) With all this in mind, we shall explain the construction of a formal integration procedure whereby $`{\displaystyle 𝖳e^{i_{(n1)ϵ}^{nϵ}\lambda ^a(t)\mathrm{\Phi }_a𝑑t}𝒟R(\lambda )}=\mathrm{EI}(\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_b\delta (\mathrm{})^2),`$ (78) and for which the integral represented by $`\mathrm{}𝒟R(\lambda )`$ is independent of the set of operators $`\{\mathrm{\Phi }_a\}`$ and the Hamiltonian operator $`(t)`$ for all $`t`$. First, introduce a formal Gaussian measure $`𝒟S_{\gamma _n}(\lambda )`$ such that $`{\displaystyle 𝖳e^{i_{(n1)ϵ}^{nϵ}\lambda ^a(t)\mathrm{\Phi }_a𝑑t}𝒟S_{\gamma _n}(\lambda )}`$ $`=𝒩{\displaystyle 𝖳e^{i_{(n1)ϵ}^{nϵ}\lambda ^a(t)\mathrm{\Phi }_a𝑑t}e^{(i/4\gamma _n)_{(n1)ϵ}^{nϵ}\lambda ^a(t)(M^1)_{ab}\lambda ^b(t)𝑑t}\mathrm{\Pi }_a𝒟\lambda ^a}`$ $`=e^{iϵ\gamma _n(\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_b)}.`$ (79) The second and last step in the construction involves an integration over $`\gamma _n`$ given by $`{\displaystyle e^{iϵ\gamma _n(\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_b)}𝑑\mathrm{\Gamma }(\gamma _n)}`$ $`\underset{\zeta 0^+}{lim}\underset{L\mathrm{}}{lim}{\displaystyle _L^L}e^{iϵ\gamma _n(\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_b)}{\displaystyle \frac{\mathrm{sin}[ϵ(\delta ^2+\zeta )\gamma _n]}{\pi \gamma _n}}𝑑\gamma _n`$ $`=\mathrm{EI}(ϵ\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_bϵ\delta ^2)`$ $`=\mathrm{EI}(\mathrm{\Phi }_aM^{ab}\mathrm{\Phi }_b\delta ^2),`$ (80) which achieves our goal. We note that if the final limit is replaced by $`lim_{\zeta 0^{}}`$, the result becomes $`\mathrm{EI}(\mathrm{\Phi }_\alpha M^{\alpha \beta }\mathrm{\Phi }_\beta <\delta ^2)`$. We normally symbolize the pair of operations by $`\mathrm{}𝒟R(\lambda )`$, leaving the integral over $`\gamma _n`$ implicit. Remark: For notational simplicity throughout this article, we generally let $`{\displaystyle e^{i\gamma X^2}\frac{\mathrm{sin}(\delta ^2\gamma )}{\pi \gamma }𝑑\gamma }`$ $`\underset{\zeta 0^+}{lim}\underset{L\mathrm{}}{lim}{\displaystyle _L^L}e^{i\gamma X^2}{\displaystyle \frac{\mathrm{sin}[(\delta ^2+\zeta )\gamma ]}{\pi \gamma }}𝑑\gamma `$ $`=\mathrm{EI}(X^2\delta ^2).`$ (81) With (80) we have found a single, universal procedure to create the regularized projection operator $`\mathrm{EI}`$ from the set of constraint operators in a manner that is completely independent of the nature of the constraints themselves. ### 4.4 Basic second-class constraint example Consider the two degree of freedom system determined by $`I=[p\dot{q}+r\dot{s}H(p,q,r,s)\lambda _1r\lambda _2s]𝑑t,`$ (82) where we have called the variables of the second degree of freedom $`r,s`$, and $`H`$ is not specified further. The coherent states satisfy $`|p,q,r,s=|p,q|r,s`$, which will be useful. We adopt (4.2) as our formal path integral in the present case, and choose $`\mathrm{EI}=e^{i(\xi _1R+\xi _2S)}e^{(\xi _1^2+\xi _2^2)/4}𝑑\xi _1𝑑\xi _2/(2\pi )`$ $`=\mathrm{EI}(R^2+S^2\mathrm{})|0_20_2|`$ (83) which is a projection operator onto the fiducial vector for the second (constrained) degree of freedom only. With this choice it follows that $`ip,q,r,s|\mathrm{EI}(d/dt)\mathrm{EI}|p,q,r,s/p,q,r,s|\mathrm{EI}|p,q,r,s`$ $`=ip,q|(d/dt)|p,q\mathrm{}(d/dt)\mathrm{ln}[0_2|r,s]`$ $`=p\dot{q}\mathrm{}(d/dt)\mathrm{ln}[0_2|r,s],`$ (84) and $`p,q,r,s|\mathrm{EI}(P,Q,R,S)\mathrm{EI}|p,q,r,s/p,q,r,s|\mathrm{EI}|p,q,r,s`$ $`=p,q,0,0|(P,Q,R,S)|p,q,0,0`$ $`=H(p,q,0,0).`$ (85) Consequently, for this example, (71) becomes $`{\displaystyle \mathrm{exp}\{i[p\dot{q}H(p,q,0,0)]𝑑t\}𝒟p𝒟q\times r^{\prime \prime },s^{\prime \prime }|0_20_2|r^{},s^{}},`$ (86) where we have used the fact that at every time slice $`r,s|\mathrm{EI}|r,s𝑑r𝑑s/(2\pi )=|0_2|r,s|^2𝑑r𝑑s/(2\pi )=1.`$ (87) Observe, in this path integral quantization, that no variables have been eliminated nor has any domain of integration been reduced; moreover, the operators $`R`$ and $`S`$ have remained unchanged. Also observe that the result in (86) is clearly a product of two distinct factors. The first factor describes the true dynamics as if we had solved for the classical constraints and substituted $`r=0`$ and $`s=0`$ in the classical action from the very beginning, while the second factor characterizes a one-dimensional Hilbert space for the second degree of freedom. Thus we can also drop the second factor completely as well as all the integrations over $`r`$ and $`s`$ and still retain the same physics. In this manner we recover the standard result without the use of Dirac brackets or having to initially eliminate the second-class constraints from the theory. ### 4.5 Conversion method One common method to treat second-class constraints is to convert them to first-class constraints and to follow the available procedures for such systems; see, e.g., . Let us first argue classically, and take as an example a single degree of freedom with canonical variables $`p`$ and $`q`$, a vanishing Hamiltonian, and the second-class constraints $`p=0`$ and $`q=0`$. This situation may be described by the classical action $`I=[p\dot{q}\lambda p\xi q]𝑑t,`$ (88) where $`\lambda `$ and $`\xi `$ denote Lagrange multipliers. Next, let us introduce a second canonical pair, say $`r`$ and $`s`$, and adopt the classical action $`I^{}=[p\dot{q}+r\dot{s}\lambda (p+r)\xi (qs)]𝑑t.`$ (89) Now the two constraints read $`p+r=0`$ and $`qs=0`$ with a Poisson bracket $`\{p+r,qs\}=0`$, characteristic of first-class constraints. We obtain the original problem by imposing the (consistent) gauge conditions that $`r=0`$ and $`s=0`$. Let us look at this example from the projection operator, coherent state approach. In the first version with one pair of variables, we are led to the reproducing kernel $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}(P^2+Q^2\mathrm{})|p^{},q^{}`$ $`=p^{\prime \prime },q^{\prime \prime }|00|p^{},q^{}`$ $`=e^{{\scriptscriptstyle \frac{1}{4}}\left(p^{\prime \prime 2}+q^{\prime \prime 2}2ip^{\prime \prime }q^{\prime \prime }\right)}e^{{\scriptscriptstyle \frac{1}{4}}\left(p^2+q^2+2ip^{}q^{}\right)},`$ (90) which provides a “bench mark” for this example. As expected the result is a one-dimensional Hilbert space. In the second version of this problem, we start with the expression $`p^{\prime \prime },q^{\prime \prime },r^{\prime \prime },s^{\prime \prime }|\mathrm{EI}((P+R)^2+(QS)^2\delta ^2)|p^{},q^{},r^{},s^{}`$ (91) which involves a constraint with zero in the continuous spectrum. Therefore, following previous examples, we multiply this expression with a suitable factor $`c_\delta `$ and take the limit as $`\delta 0`$. This factor can be chosen so that $`\underset{\delta 0}{lim}c_\delta p^{\prime \prime },q^{\prime \prime },r^{\prime \prime },s^{\prime \prime }|\mathrm{EI}((P+R)^2+(QS)^2\delta ^2)|p^{},q^{},r^{},s^{}`$ $`=e^{{\scriptscriptstyle \frac{1}{4}}\left[\left(p^{\prime \prime }+r^{\prime \prime }\right)^2+\left(q^{\prime \prime }s^{\prime \prime }\right)^2\right]+{\scriptscriptstyle \frac{1}{2}}i\left(p^{\prime \prime }r^{\prime \prime }\right)\left(q^{\prime \prime }s^{\prime \prime }\right)}`$ $`\times e^{{\scriptscriptstyle \frac{1}{2}}i\left(p^{}r^{}\right)\left(q^{}s^{}\right){\scriptscriptstyle \frac{1}{4}}\left[\left(p^{}+r^{}\right)^2+\left(q^{}s^{}\right)^2\right]},`$ (92) an expression which also describes a one-dimensional Hilbert space. This is a different (but equivalent) representation for the one-dimensional Hilbert space than the one found above. Since it is only one-dimensional we can reduce this reproducing kernel even further, in the fashion illustrated earlier, by choosing a “gauge” where $`r^{\prime \prime }=s^{\prime \prime }=r^{}=s^{}=0`$. When this is done the result becomes $`e^{{\scriptscriptstyle \frac{1}{4}}\left(p^{\prime \prime 2}+q^{\prime \prime 2}2ip^{\prime \prime }q^{\prime \prime }\right)}e^{{\scriptscriptstyle \frac{1}{4}}\left(p^2+q^2+2ip^{}q^{}\right)},`$ (93) which is identical to the expression (90) found by quantization of the second-class constraints directly. In this manner we see how the conversion method, in which second-class constraints are turned into first-class constraints by the introduction of auxiliary degrees of freedom, appears within the projection operator, coherent state approach as well. Applications of the conversion method made within the projection operator approach may be found in . ### 4.6 Equivalent representations In dealing with quantum mechanics, one may employ many different—yet equivalent—representations of the vectors and operators involved. While, in certain circumstances, some representations may be more convenient than others, the notion that some representations are “better” than others should be resisted. In the context of coherent-state representations, for example, a change of the fiducial vector leads to an equivalent representation. If, for a rather general (normalized) fiducial vector $`|\eta `$, we set $`|p,q;\eta e^{iqP}e^{ipQ}|\eta ,`$ (94) then $`\psi (p,q;\eta )p,q;\eta |\psi `$ (95) defines $`\eta `$-dependent representatives of the abstract vector $`|\psi `$. However, all representation-dependent aspects disappear when physical questions are asked such as $`|\psi (p,q;\eta )|^2(dpdq/2\pi )=\psi |\psi .`$ (96) More general representation issues may be addressed by using arbitrary unitary operators, say $`V`$. Thus if $`|p,q`$ denotes elements of one (say) coherent state basis, then $`|p,q;VV^{}|p,q`$ denotes the elements of another basis. Vector and operator representatives, $`\psi (p,q;V)p,q;V|\psi `$ and $`A(p^{},q^{};V:p,q;V)p^{},q^{};V|𝒜|p,q;V`$, respectively, provide equivalent sets of functional representatives for different $`V`$. Evidently the physics is unchanged in this transformation; only the intermediate mathematical representatives are affected. This formulation is similar to passive coordinate transformations in other disciplines. Another version similar to active coordinate transformations is also possible. In this version the basis vectors, say $`|p,q`$, for all relevant $`(p,q)`$, remain unchanged; instead, the abstract vectors $`|\psi `$ and operators $`𝒜`$, etc., are transformed: $`|\psi V|\psi `$, $`𝒜V𝒜V^{}`$, etc. It is this form of equivalence that we turn to next. ### 4.7 Equivalence of criteria for <br>second-class constraints Let us return to the simple example of second-class constraints discussed above where, classically, $`p=q=0`$. In the associated quantum theory, we chose to express these constraints with the help of the projection operator $`\mathrm{EI}=\mathrm{EI}(P^2+Q^2\mathrm{})=|00|`$, namely, the projection operator onto the ground state of the “Hamiltonian” $`P^2+Q^2`$. In turn, this expression led directly to the coherent-state representation of $`\mathrm{EI}`$ given by $`p^{},q^{}|\mathrm{EI}|p,q=p^{},q^{}|00|p,q`$. However, the question arises, what is special about the combination $`P^2+Q^2`$? As we shall now argue, any other possible choice leads to an equivalent representation. As a first example, consider $`\mathrm{EI}(P^2+\omega ^2Q^2\omega \mathrm{})=|0;\omega 0;\omega |=V_\omega ^{}|00|V_\omega ,`$ (97) where $`V_\omega `$ denotes a suitable unitary operator, which establishes the equivalence for any $`\omega `$, $`0<\omega <\mathrm{}`$. We emphasize that we do not assert the unitary equivalence of $`P^2+Q^2`$ and $`P^2+\omega ^2Q^2`$ for any value of $`\omega 1`$, only that $`|0;\omega `$ and $`|0`$ are unitarily related—as are any two unit vectors in Hilbert space. Furthermore, there is nothing sacred about the quadratic combination. For example, for any $`0<\lambda <\mathrm{}`$, consider $`\mathrm{EI}(P^2+\lambda Q^4\delta (\mathrm{})^2)|0,\lambda 0,\lambda |`$, where we have adjusted $`\delta (\mathrm{})`$ to the lowest eigenvalue so as to include only a single eigenvector, $`|0,\lambda `$. Since there exists a unitary operator $`V_\lambda `$ such that $`0,\lambda |=0|V_\lambda `$, this choice of projection operator leads to an equivalent coherent-state representation as well. More generally, we are led to reconsider the projection operator $`\mathrm{EI}(\mathrm{\Sigma }_a\mathrm{\Phi }_a^2\delta (\mathrm{})^2)={\displaystyle \underset{j=1}{\overset{J}{}}}|jj|,`$ (98) where $`j|k=\delta _{jk}`$ and $`1J\mathrm{}`$, as determined by the choice of $`\delta (\mathrm{})`$. Since all $`J`$-dimensional subspaces are unitarily equivalent to each other (with suitable care taken when $`J=\mathrm{}`$), the given prescription is entirely equivalent to any other version, such as $`\mathrm{EI}((\mathrm{\Phi }_a)\stackrel{~}{\delta }(\mathrm{})^2)={\displaystyle \underset{𝗃=1}{\overset{𝖩}{}}}|𝗃𝗃|,`$ (99) where $`𝗃|𝗄=\delta _{\mathrm{𝗃𝗄}}`$, provided that $`\stackrel{~}{\delta }(\mathrm{})`$ may be—and is—chosen so that $`𝖩=J`$. Here $`(\mathrm{\Phi }_a)`$ denotes a nonnegative self-adjoint operator that includes all the constraint operators, and for very small $`\stackrel{~}{\delta }(\mathrm{})^2`$ forces the spectral contribution of each constraint operator to be correspondingly small, just as is the case in (98). In summary, the general, quadratic criterion we have adopted in (98) has been chosen for simplicity and convenience; any other restriction on the constraint operators leads to an equivalent theory, as in (99), provided that the dimensionality of $`\mathrm{EI}`$ remains the same. ## 5 SELECTED EXAMPLES OF <br>FIRST-CLASS CONSTRAINTS ### 5.1 General configuration space geometry Although we shall discuss constraints that lead to a general configuration space geometry in this section, we shall for the most part use rather simple illustrative examples. To begin with let us consider the constraint $`{\displaystyle \underset{j=1}{\overset{J}{}}}(q^j)^2=1,`$ (100) a condition which puts the classical problem on a (hyper)sphere of unit radius. For convenience in what follows we shall focus as well on the case of a vanishing Hamiltonian so as to isolate clearly the consequences of the constraint independently of any dynamical effects. Adopting a standard vector inner product notation and a different kinematic term, consider the formal path integral $`{\displaystyle \mathrm{exp}\{i[q\dot{p}\lambda (q^21)]𝑑t\}𝒟p𝒟q𝒟C(\lambda )},`$ (101) the result of which is given by $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$ (102) where $`\mathrm{EI}={\displaystyle _{\mathrm{}}^{\mathrm{}}}e^{i\lambda (Q^21)}{\displaystyle \frac{\mathrm{sin}(\delta \lambda )}{\pi \lambda }}𝑑\lambda =\mathrm{EI}(\delta Q^21\delta ).`$ (103) In order, ultimately, to obtain a suitable reduction of the reproducing kernel in the present case, we allow for fiducial vectors other than harmonic oscillator ground states. Thus we let $`|\eta `$ denote a general unit vector for the moment; its required properties will emerge from our analysis. In accordance with (101), we choose a phase convention for the coherent states—in particular, in (18) we multiply by $`e^{ipq}`$—so that now the Schrödinger representation of the coherent states reads $`x|p,q=e^{ipx}\eta (xq),`$ (104) which leads immediately to the expression $`p^{\prime \prime },q^{\prime \prime }|p^{},q^{}={\displaystyle \eta ^{}(xq^{\prime \prime })e^{i(p^{\prime \prime }p^{})x}\eta (xq^{})d^Jx}.`$ (105) Consequently, the reproducing kernel that incorporates the projection operator is given, for $`0<\delta <1`$, by $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}={\displaystyle _{1\delta x^21+\delta }}\eta ^{}(xq^{\prime \prime })e^{i(p^{\prime \prime }p^{})x}\eta (xq^{})d^Jx.`$ (106) Since $`\mathrm{EI}`$ represents a projection operator, it is evident that this expression defines a reproducing kernel which admits a local integral for its inner product (for any normalized $`\eta `$) with a measure $`d^Jpd^Jq/(2\pi )^J`$ and an integration domain $`\mathrm{IR}^{2J}`$. However, if we are willing to restrict our choice of fiducial vector, we can reduce the number of integration variables and change the domain of integration in a meaningful way. Recall that the group $`\mathrm{E}(J)`$, the Euclidean group in $`J`$-dimensions, consists of rotations that preserve the unit (hyper)sphere in $`J`$-dimensions, as well as $`J`$ translations. As emphasized by Isham , this is the natural canonical group for a system confined to the surface of a (hyper)sphere in $`J`$ dimensions. We can adapt our present coherent states to be coherent states for the group $`\mathrm{E}(J)`$ without difficulty. To that end consider the reduction of the reproducing kernel (106) to one for which $`q^{\prime \prime 2}=q^21`$. To illustrate the process as clearly as possible let us choose $`J=2`$. As a consequence we introduce $`a^{\prime \prime },b^{\prime \prime },c^{\prime \prime }|a^{},b^{},c^{}p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}_{q^{\prime \prime 2}=q^2=1},`$ (107) where $`ap_1`$, $`bp_2`$, and $`c`$ arises from the identification $`q^1\mathrm{cos}(c)`$ and $`q^2\mathrm{sin}(c)`$, all relations holding for both end points. Expressed in terms of polar coordinates, $`r,\varphi `$, the reduced reproducing kernel becomes $`a^{\prime \prime },b^{\prime \prime },c^{\prime \prime }|a^{},b^{},c^{}`$ $`={\displaystyle _{|r^21|\delta }}\eta ^{}(r,\varphi c^{\prime \prime })e^{i(a^{\prime \prime }a^{})r\mathrm{cos}\varphi i(b^{\prime \prime }b^{})r\mathrm{sin}\varphi }\eta (r,\varphi c^{})r𝑑r𝑑\varphi .`$ (108) We next seek to choose $`\eta `$, if at all possible, in such a way that the inner product of this new (reduced) reproducing kernel admits a local integral for its inner product. As a starting point we choose the left-invariant group measure for $`\mathrm{E}(2)`$ which is given by $`Mdadbdc`$, $`M`$ a constant, with an integration domain $`\mathrm{IR}^2\times S^1`$. Therefore, we are led to propose that $`{\displaystyle _{|r^21|<\delta }\eta ^{}(r,\varphi c^{\prime \prime })e^{i(a^{\prime \prime }a)r\mathrm{cos}\varphi i(b^{\prime \prime }b)r\mathrm{sin}\varphi }\eta (r,\varphi c)r𝑑r𝑑\varphi }`$ $`\times {\displaystyle _{|\rho ^21|<\delta }}\eta ^{}(\rho ,\theta c)e^{i(aa^{})\rho \mathrm{cos}\theta i(bb^{})\rho \mathrm{sin}\theta }\eta (\rho ,\theta c^{})\rho d\rho d\theta `$ $`\times Mdadbdc`$ $`=(2\pi )^2M{\displaystyle \eta ^{}(r,\varphi c^{\prime \prime })e^{i(a^{\prime \prime }a^{})r\mathrm{cos}\varphi i(b^{\prime \prime }b^{})r\mathrm{sin}\varphi }\eta (r,\varphi c^{})r𝑑r𝑑\varphi }`$ $`\times {\displaystyle }|\eta (r,c)|^2dc,`$ (109) which leads to the desired result provided (i) $`{\displaystyle _0^{2\pi }}|\eta (r,c)|^2𝑑c=P,0<P<\mathrm{},`$ (110) is independent of $`r`$, $`|r^21|<\delta `$, and (ii) $`M=[(2\pi )^2P]^1`$. Given a general nonvanishing vector $`\xi (r,\varphi )`$, a vector satisfying (110) may always be given by $`\eta (r,\varphi )=\xi (r,\varphi )/\sqrt{_0^{2\pi }|\xi (r,\theta )|^2𝑑\theta }`$ (111) provided the denominator is positive, and which specifically leads to $`P=1`$. In this way we have reproduced the $`\mathrm{E}(2)`$-coherent states of Ref. , even including the necessity for a small interval of integration in $`r`$, and where fiducial vectors satisfying (110) were called “surface constant”. Dynamics consistent with the constraint $`q^2=1`$ is obtained in the $`\mathrm{E}(2)`$ case by choosing a Hamiltonian that is a function of the coordinates on the circle, namely $`\mathrm{cos}(\theta )`$ and $`\mathrm{sin}(\theta )`$, as well as the rotation generator of $`\mathrm{E}(2)`$, i.e., $`i/\theta `$. We refer the reader to for a further discussion of $`\mathrm{E}(2)`$-coherent states as well as a discussion of the introduction of compatible dynamics. An analogous discussion can be given for the classical constraint $`q^2=1`$ for any value of $`J>2`$. Not only can compact (hyper)spherical configuration spaces be treated in this way, but one may also treat noncompact (hyper)pseudospherical spaces defined by the constraint $`\mathrm{\Sigma }_{i=1}^Iq^{i\mathrm{\hspace{0.17em}2}}\mathrm{\Sigma }_{j=I+1}^Jq^{j\mathrm{\hspace{0.17em}2}}=1,1IJ1,`$ (112) appropriate to the Euclidean group $`\mathrm{E}(I,JI)`$. Such an analysis would lead to $`\mathrm{E}(I,JI)`$-coherent states. Finally, we comment on the constraint of a general curved configuration space which can be defined by a set of compatible constraints $`\varphi _a(q)=0`$. Clearly these constraints satisfy $`\{\varphi _a(q),\varphi _b(q)\}=0`$, and define a $`(JA)`$-dimensional configuration space in the original Euclidean configuration space $`\mathrm{IR}^J`$. The relevant projection operator $`\mathrm{EI}=\mathrm{EI}(\mathrm{\Sigma }\mathrm{\Phi }_a^2(Q)\delta ^2)`$ is defined in an evident fashion, and the reproducing kernel incorporating the projection operator is defined in analogy with the prior discussion. This reproducing kernel enjoys a local integral representation for its inner product, in fact, this integral is with the same measure and integration domain as without the projection operator. What differs in the present case is that when the reproducing kernel is put on the constraint manifold, the resultant coherent states are generally not defined by the action of a group on a fixed fiducial vector. In short, the relevant coherent states are not group generated, which, in fact, is consistent with their most basic definition; see, e.g., . ### 5.2 Finite-dimensional Hilbert space examples Let us consider the case of two degrees of freedom with a “classical” action function given by $`I=[\frac{1}{2}(p_1\dot{q}_1q_1\dot{p}_1+p_2\dot{q}_2q_2\dot{p}_2)\lambda (p_1^2+p_2^2+q_1^2+q_2^24s\mathrm{})]𝑑t`$ (113) For clarity of presentation, we explicitly include $`\mathrm{}`$ in our classical action, and we continue to make it explicit it throughout this section. With the present phase convention for the coherent states, the unconstrained reproducing kernel is given by $`p^{\prime \prime },q^{\prime \prime }|p^{},q^{}z^{\prime \prime }|z^{}`$ $`=\mathrm{exp}[\mathrm{\Sigma }_{j=1}^2(\frac{1}{2}|z_j^{\prime \prime }|^2+z_{}^{\prime \prime }{}_{j}{}^{}z_j^{}\frac{1}{2}|z_j^{}|^2)]`$ (114) where $`z_j(q_j+ip_j)/\sqrt{2\mathrm{}}`$ for each of the end points. We next observe that the constraint operator $`\mathrm{\Phi }=:P_1^2+P_2^2+Q_1^2+Q_2^2:4s\mathrm{}11`$ (115) has discrete eigenvalues, i.e., $`2(n_1+n_22s)\mathrm{}`$, where $`n_1`$ and $`n_2`$ are nonnegative integers, based on the choice of $`|\eta `$ as the ground state for each oscillator. To satisfy $`\mathrm{\Phi }=0`$ it is necessary that $`2s`$ be an integer in which case the quantum constraint subspace is $`(2s+1)`$-dimensional. The projection operator in the present case is defined by $`\mathrm{EI}=\pi ^1{\displaystyle _0^\pi }\mathrm{exp}[i\lambda (:P_1^2+P_2^2+Q_1^2+Q_2^2:4s\mathrm{}11)/\mathrm{}]d\lambda `$ (116) which projects onto the appropriate $`(2s+1)`$-dimensional subspace. It is straightforward to demonstrate that $`z^{\prime \prime }|\mathrm{EI}|z^{}=\mathrm{exp}[\frac{1}{2}\mathrm{\Sigma }_{j=1}^2(|z_j^{\prime \prime }|^2+|z_j^{}|^2)][(2s)!]^1(z_{}^{\prime \prime }{}_{1}{}^{}z_1^{}+z_{}^{\prime \prime }{}_{2}{}^{}z_2^{})^{2s}`$ $`=\mathrm{exp}[\frac{1}{2}\mathrm{\Sigma }_{j=1}^2(|z_j^{\prime \prime }|^2+|z_j^{}|^2)]_{k=0}^{2s}[k!(2sk)!]^1(z_{}^{\prime \prime }{}_{1}{}^{}z_1^{})^k(z_{}^{\prime \prime }{}_{2}{}^{}z_2^{})^{2sk}`$ (117) The projected reproducing kernel in this case corresponds to a finite dimensional Hilbert space; nevertheless, the inner product is given by the same measure and integration domain as in the original, unprojected, infinite dimensional Hilbert space! Of course, there are other, simpler and more familiar ways to represent a finite-dimensional Hilbert space; but any other representation is evidently equivalent to the one described here. As the notation suggests the present quantum constraint subspace provides a natural carrier space for an irreducible representation of $`\mathrm{SU}(2)`$ with spin $`s`$. We observe that the following three expressions represent generators of the classical rotation group in their action on the constraint hypersurface: $`s_x=\frac{1}{2}(p_1p_2+q_1q_2),`$ $`s_y=\frac{1}{2}(q_1p_2p_1q_2),`$ $`s_z=\frac{1}{4}(p_1^2+q_1^2p_2^2q_2^2).`$ (118) Thus these quantities serve as potential ingredients for a Hamiltonian which is compatible with the constraint. Although not the subject of this section, we may also observe that an analogous discussion holds in case of the constraint $`\varphi (p,q)=p_1^2+q_1^2p_2^2q_2^22k\mathrm{}=0,`$ (119) where $`k`$ is an integer, and the resultant reduced Hilbert space is infinite dimensional for any integral $`k`$ value. In this case the relevant group is SU(1,1). ### 5.3 Helix model In , Friedberg, Lee, Pang, and Ren analyzed the so-called helix model. For details of this model (see also ) and its possible role as a simple analogue of the Gribov problem in non-Abelian gauge models, we refer the reader to their paper. We begin with the classical Hamiltonian for a three-degree of freedom system given by $`H=\frac{1}{2}(p_1^2+p_2^2+p_3^2)+U(q_1^2+q_2^2)+\lambda [g(p_2q_1q_2p_1)+p_3],`$ (120) where $`U`$ denotes the potential, which hereafter, following , we shall choose as harmonic, namely $`U(q_1^2+q_2^2)=\omega ^2(q_1^2+q_2^2)/2`$ , because then this special model is fully soluble. Here, $`g>0`$ is a coupling constant, and $`\lambda =\lambda (t)`$ is the Lagrange multiplier which enforces the single first-class constraint $`\varphi (p,q)=g(p_2q_1q_2p_1)+p_3=0.`$ (121) For the first two degrees of freedom we choose coherent states with the phase convention adopted for the previous example, while for the third degree of freedom we return to the original phase convention. This choice means that we consider the formal coherent state path integral given by $`{\displaystyle }\mathrm{exp}((i\{\frac{1}{2}(p_1\dot{q}_1q_1\dot{p}_1)+\frac{1}{2}(p_2\dot{q}_2q_2\dot{p}_2)+p_3\dot{q}_3`$ $`\frac{1}{2}(p_1^2+p_2^2+p_3^2)\frac{1}{2}\omega ^2(q_1^2+q_2^2)`$ $`\lambda [g(p_2q_1q_2p_1)+p_3]\}dt))𝒟\mu (p,q)𝒟C(\lambda )`$ $`=z_1^{\prime \prime },z_2^{\prime \prime },p_3^{\prime \prime },q_3^{\prime \prime }|e^{iT}\mathrm{EI}|z_1^{},z_2^{},p_3^{},q_3^{}.`$ (122) In the present case the relevant projection operator $`\mathrm{EI}`$ is given (for $`\mathrm{}=1`$, and $`0<\delta g`$) by $`\mathrm{EI}=\mathrm{EI}((gL_3+P_3)^2\delta ^2)={\displaystyle \underset{m=\mathrm{}}{\overset{\mathrm{}}{}}}\mathrm{EI}((gm+P_3)^2\delta ^2)\mathrm{EI}(L_3=m),`$ (123) where we have used the familiar spectrum for the rotation generator $`L_3`$. If $`_0`$ denotes the harmonic oscillator Hamiltonian for the first two degrees of freedom, then it follows that $`z_1^{\prime \prime },z_2^{\prime \prime },p_3^{\prime \prime },q_3^{\prime \prime }|e^{iT}\mathrm{EI}|z_1^{},z_2^{},p_3^{},q_3^{}`$ $`={\displaystyle \underset{m=\mathrm{}}{\overset{\mathrm{}}{}}}z_1^{\prime \prime },z_2^{\prime \prime }|e^{i_0T}\mathrm{EI}(L_3=m)|z_1^{},z_2^{}`$ $`\times p_3^{\prime \prime },q_3^{\prime \prime }|e^{iP_3^2T/2}\mathrm{EI}(\delta gm+P_3\delta )|p_3^{},q_3^{}`$ $`=\mathrm{exp}[\frac{1}{2}(|z_1^{\prime \prime }|^2+|z_2^{\prime \prime }|^2+|z_1^{}|^2+|z_2^{}|^2)]`$ $`\times {\displaystyle \underset{m=\mathrm{}}{\overset{\mathrm{}}{}}}\left\{{\displaystyle \frac{(z_1^{\prime \prime }+iz_2^{\prime \prime })(z_1^{}iz_2^{})}{(z^{\prime \prime }iz_2^{\prime \prime })(z_1^{}+iz_2^{})}}\right\}^{m/2}I_m(\sqrt{(z_1^{\prime \prime 2}+z_2^{\prime \prime 2})(z_1^2+z_2^2)}e^{i\omega T})`$ $`\times \mathrm{exp}[\frac{1}{2}(gm+p_3^{\prime \prime })^2\frac{1}{2}(gm+p_3^{})^2i\frac{1}{2}g^2m^2Tigm(q_3^{\prime \prime }q_3^{})]`$ $`\times {\displaystyle \frac{2}{\sqrt{\pi }}}{\displaystyle \frac{\mathrm{sin}[\delta (q_3^{\prime \prime }q_3^{})]}{(q_3^{\prime \prime }q_3^{})}}+O(\delta ^2),`$ (124) where $`I_m`$ denotes the usual Bessel function. We observe that the spectrum for the Hamiltonian agrees with the results of Ref. , and moreover, to leading order in $`\delta `$, we have obtained gauge-invariant results, i.e., insensitivity to any choice of the Lagrange multiplier function $`\lambda (t)`$, merely by projecting onto the quantum constraint subspace at $`t=0`$. The constrained propagator (124) is composed with the same measure and integration domain as is the unconstrained propagator. We may also divide the constrained propagator by $`\delta `$ and take the limit $`\delta 0`$. The result is a new functional expression for the propagator that fully satisfies the constraint condition, but one that no longer admits an inner product with the same measure and integration domain as before. ### 5.4 Reparameterization invariant dynamics Let us start with a single degree of freedom $`(J=1)`$ and the action $`[p\dot{q}H(p,q)]𝑑t.`$ (125) We next promote the independent variable $`t`$ to a dynamical variable, introduce $`s`$ as its conjugate momentum (often called $`p_t`$), enforce the constraint $`s+H(p,q)=0`$, and lastly introduce $`\tau `$ as a new independent variable. This modification is realized by means of the classical action $`\{pq^{}+st^{}\lambda [s+H(p,q)]\}𝑑\tau ,`$ (126) where $`q^{}=dq/d\tau `$, $`t^{}=dt/d\tau `$, and $`\lambda =\lambda (\tau )`$ is a Lagrange multiplier. The coherent-state path integral is constructed so that $`{\displaystyle \mathrm{exp}((i\{pq^{}+st^{}\lambda [s+H(p,q)]\}𝑑t))𝒟p𝒟q𝒟s𝒟t𝒟C(\lambda )}`$ $`=p^{\prime \prime },q^{\prime \prime },s^{\prime \prime },t^{\prime \prime }|\mathrm{EI}|p^{},q^{},s^{},t^{},`$ (127) where $`\mathrm{EI}={\displaystyle _{\mathrm{}}^{\mathrm{}}}e^{i\xi [S+(P,Q)]}{\displaystyle \frac{\mathrm{sin}(\delta \xi )}{\pi \xi }}𝑑\xi `$ $`=\mathrm{EI}(\delta S+(P,Q)\delta ).`$ (128) The result in (127) and (128) represents as far as we can go without choosing $`(P,Q)`$. To gain further insight into such expressions, we specialize to the case of the nonrelativistic free particle, $`=P^2/2`$. Then it follows that $`p^{\prime \prime },q^{\prime \prime },s^{\prime \prime },t^{\prime \prime }|\mathrm{EI}|p^{},q^{},s^{},t^{}`$ $`=\pi ^1{\displaystyle _{\mathrm{}}^{\mathrm{}}}\mathrm{exp}[\frac{1}{2}(kp^{\prime \prime })^2\frac{1}{2}(\frac{1}{2}k^2+s^{\prime \prime })^2`$ $`+ik(q^{\prime \prime }q^{})i\frac{1}{2}k^2(t^{\prime \prime }t^{})`$ $`\frac{1}{2}(kp^{})^2\frac{1}{2}(\frac{1}{2}k^2+s^{})^2]dk`$ $`\times {\displaystyle \frac{2\mathrm{sin}[\delta (t^{\prime \prime }t^{})]}{(t^{\prime \prime }t^{})}}+O(\delta ^2).`$ (129) For any $`\delta `$ such that $`0<\delta 1`$, we observe that this expression represents a reproducing kernel which in turn defines an associated reproducing kernel Hilbert space composed, as usual, of bounded, continuous functions given, for arbitrary complex numbers $`\{\alpha _k\}`$, phase-space points $`\{p_k,q_k,s_k,t_k\}`$, and $`K<\mathrm{}`$, by $`\psi (p,q,s,t){\displaystyle \underset{k=0}{\overset{K}{}}}\alpha _kp,q,s,t|\mathrm{EI}|p_k,q_k,s_k,t_k,`$ (130) or as the limit of Cauchy sequences of such functions in the norm defined by means of the inner product given by $`(\psi ,\psi )=|\psi (p,q,s,t)|^2𝑑p𝑑q𝑑s𝑑t/(2\pi )^2`$ (131) integrated over $`\mathrm{IR}^4`$. Let us next consider the reduction of the reproducing kernel given by $`p^{\prime \prime },q^{\prime \prime },t^{\prime \prime }|p^{},q^{},t^{}`$ $`\underset{\delta 0}{lim}{\displaystyle \frac{1}{4\sqrt{\pi }\delta }}{\displaystyle p^{\prime \prime },q^{\prime \prime },s^{\prime \prime },t^{\prime \prime }|\mathrm{EI}|p^{},q^{},s^{},t^{}𝑑s^{\prime \prime }𝑑s^{}}`$ $`=\pi ^{1/2}{\displaystyle }\mathrm{exp}[\frac{1}{2}(kp^{\prime \prime })^2\frac{1}{2}(kp^{})^2`$ $`+ik(q^{\prime \prime }q^{})i\frac{1}{2}k^2(t^{\prime \prime }t^{})]dk,`$ (132) which in turn generates a new reproducing kernel in the indicated variables. For the resultant kernel it is straightforward to demonstrate, for any $`t`$, that $`{\displaystyle p^{\prime \prime },q^{\prime \prime },t^{\prime \prime }|p,q,tp,q,t|p^{},q^{},t^{}𝑑p𝑑q/(2\pi )}=p^{\prime \prime },q^{\prime \prime },t^{\prime \prime }|p^{},q^{},t^{}.`$ (133) This relation implies that the span of the vectors $`\{|p,q|p,q,0\}`$ is identical with the span of the vectors $`\{|p,q,t\}`$, meaning further that the states $`\{|p,q,t\}`$ form a set of extended coherent states, which are “extended” with respect to $`t`$ in the sense of Ref. . Observe how the time variable has become distinguished by the criterion (133). Consequently, we may properly interpret $`p^{\prime \prime },q^{\prime \prime },t^{\prime \prime }|p^{},q^{},t^{}p^{\prime \prime },q^{\prime \prime }|e^{i(P^2/2)(t^{\prime \prime }t^{})}|p^{},q^{},`$ (134) namely, as the conventional, single degree of freedom, coherent-state matrix element of the evolution operator appropriate to the free particle. To further demonstrate this interpretation as the dynamics of the free particle, we may pass to sharp $`q`$ matrix elements with the observation that $`q^{\prime \prime }|e^{i(P^2/2)(t^{\prime \prime }t^{})}|q^{}`$ $`{\displaystyle \frac{\pi ^{1/2}}{(2\pi )^2}}{\displaystyle p^{\prime \prime },q^{\prime \prime }|e^{i(P^2/2)(t^{\prime \prime }t^{})}|p^{},q^{}𝑑p^{\prime \prime }𝑑p^{}}`$ $`={\displaystyle \frac{1}{2\pi }}{\displaystyle \mathrm{exp}[ik(q^{\prime \prime }q^{})i\frac{1}{2}k^2(t^{\prime \prime }t^{})]𝑑k}`$ $`={\displaystyle \frac{e^{i(q^{\prime \prime }q^{})^2/2(t^{\prime \prime }t^{})}}{\sqrt{2\pi i(t^{\prime \prime }t^{})}}},`$ (135) which is clearly the usual result. ### 5.5 Elevating the Lagrange multiplier to an <br>additional dynamical variable Sometimes it is useful to consider an alternative formulation of a system with constraints in which the initial Lagrange multipliers are regarded as dynamical variables, complete with their own conjugate variables, and to introduce new constraints as needed. For example, let us start with a single degree of freedom system with a single first-class constraint specified by the action functional $`[p\dot{q}H(p,q)\lambda \varphi (p,q)]𝑑t,`$ (136) where $`\varphi (p,q)`$ represents the constraint and $`\lambda `$ the Lagrange multiplier. Instead, let us replace this action functional by $`[p\dot{q}+\pi \dot{\lambda }H(p,q)\sigma \pi \theta \varphi (p,q)]𝑑t.`$ (137) In this expression we have introduced $`\pi `$ as the canonical conjugate to $`\lambda `$, the Lagrange multiplier $`\sigma `$ to enforce the constraint $`\pi =0`$, and the Lagrange multiplier $`\theta `$ to enforce the original constraint $`\varphi =0`$. Observe that $`\{\pi ,\varphi (p,q)\}=0`$, and therefore the constraints remain first class in the new form. The path integral expression for the extended form reads $`{\displaystyle \mathrm{exp}\{i[p\dot{q}+\pi \dot{\lambda }H(p,q)\sigma \pi \theta \varphi (p,q)]𝑑t\}𝒟p𝒟q𝒟\pi 𝒟\lambda 𝒟C(\sigma ,\theta )}`$ $`=p^{\prime \prime },q^{\prime \prime },\pi ^{\prime \prime },\lambda ^{\prime \prime }|e^{iT}\mathrm{EI}|p^{},q^{},\pi ^{},\lambda ^{}.`$ (138) In this expression, we may choose $`\mathrm{EI}=\mathrm{EI}(\mathrm{\Phi }(P,Q)^2\delta ^2)\mathrm{EI}(\mathrm{\Pi }^2\delta ^2)`$ (139) involving two possibly distinct reglarization parameters. Consequently, the complete propagator factors into two terms, $`p^{\prime \prime },q^{\prime \prime },\pi ^{\prime \prime },\lambda ^{\prime \prime }|e^{iT}\mathrm{EI}|p^{},q^{},\pi ^{},\lambda ^{}`$ $`=p^{\prime \prime },q^{\prime \prime }|e^{iT}\mathrm{EI}(\mathrm{\Phi }(P,Q)^2\delta ^2)|p^{},q^{}`$ $`\times \pi ^{\prime \prime },\lambda ^{\prime \prime }|\mathrm{EI}(\mathrm{\Pi }^2\delta ^2)|\pi ^{},\lambda ^{}.`$ (140) The first factor is exactly what would be found by the appropriate path integral of the original classical system with only the single constraint $`\varphi (p,q)=0`$ and the single Lagrange multiplier $`\lambda `$. The second factor represents the modification introduced by considering the extended system. Note, however, that with a suitable $`\delta ^{}`$-limit the second factor reduces to a product of terms, one depending on the “$`^{\prime \prime }`$” arguments, the other depending on the “$`^{}`$” arguments, just as was the case previously. This result for the second factor implies that it has become the reproducing kernel for a one-dimensional Hilbert space, and when multiplied by the first factor it may be ignored entirely. In this way it is found that the quantization of the original and extended systems leads to identical results. ## 6 SPECIAL APPLICATIONS ### 6.1 Algebraically inequivalent constraints The following example is suggested by Problem 5.1 in Ref. . Consider the two-degree of freedom system with vanishing Hamiltonian described by the classical action $`I=(p_1\dot{q}_1+p_2\dot{q}_2\lambda _1p_1\lambda _2p_2)𝑑t.`$ (141) The equations of motion become $`\dot{q}_j=\lambda _j,\dot{p}_j=0,p_j=0,j=1,2.`$ (142) Evidently the Poisson bracket $`\{p_1,p_2\}=0`$. As a second version of the same dynamics, consider the classical action $`I=(p_1\dot{q}_1+p_2\dot{q}_2\lambda _1p_1\lambda _2e^{cq_1}p_2)𝑑t,`$ (143) which leads to the equations of motion $`\dot{q}_1=\lambda _1,\dot{q}_2=\lambda _2e^{cq_1},\dot{p}_1=c\lambda _2e^{cq_1}p_2,\dot{p}_2=0,p_1=e^{cq_1}p_2=0`$ (144) Since $`e^{cq_1}p_2=0`$ implies that $`p_2=0`$, it follows that the two formulations are equivalent despite the fact that in the second case $`\{p_1,e^{cq_1}p_2\}=ce^{cq_1}p_2`$, which has a fundamentally different algebraic structure when $`c0`$ as compared to $`c=0`$. Let us discuss these two examples from the point of view of a coherent state, projection operator quantization. For the first version we consider $`{\displaystyle \mathrm{exp}[i(p_1\dot{q}_1+p_2\dot{q}_2\lambda _1p_1\lambda _2p_2)𝑑t]𝒟p𝒟q𝒟C(\lambda )},`$ (145) defined in a fashion to yield $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$ (146) where, for ease of evaluation, we may choose $`\mathrm{EI}=\mathrm{EI}(P_1^2\delta ^2)\mathrm{EI}(P_2^2\delta ^2).`$ (147) In particular this choice leads to the fact that $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}`$ $`=\pi ^1{\displaystyle \underset{l=1}{\overset{2}{}}}{\displaystyle _\delta ^\delta }\mathrm{exp}[\frac{1}{2}(k_lp_l^{\prime \prime })^2+ik_l(q_l^{\prime \prime }q_l^{})\frac{1}{2}(k_lp_l^{})^2]𝑑k_l.`$ (148) Let us reduce this reproducing kernel, in particular, by multiplying this expression by $`\pi /(2\delta )^2`$ and passing to the limit $`\delta 0`$. The result is the reduced reproducing kernel given by $`\mathrm{exp}[\frac{1}{2}(p_1^{\prime \prime 2}+p_2^{\prime \prime 2})]\mathrm{exp}[\frac{1}{2}(p_1^2+p_2^2)],`$ (149) which clearly characterizes a particular representation of a one-dimensional Hilbert space in which every vector is proportional to $`\mathrm{exp}[\frac{1}{2}(p_1^2+p_2^2)]`$. This example, of course, is related to the reduction examples given earlier. Moreover, we can introduce an integral representation over the remaining $`p`$ variables for the inner product if we so desire. Let us now turn attention to the second formulation of the problem by focussing \[for a different $`C(\lambda )`$\] on $`{\displaystyle \mathrm{exp}[i(p_1\dot{q}_1+p_2\dot{q}_2\lambda _1p_1\lambda _2e^{cq_1}p_2)𝑑t]𝒟p𝒟q𝒟C(\lambda )}.`$ (150) This expression again leads (for a different $`\mathrm{EI}`$) to $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{},`$ (151) where in the present case the fully reduced form of this expression is proportional to $`{\displaystyle \mathrm{exp}[\frac{1}{2}(k_2p_2^{\prime \prime })^2+ik_2(q_2^{\prime \prime }q_2^{})\frac{1}{2}(k_2p_2^{})^2]}`$ $`\times \mathrm{exp}[\frac{1}{2}(k_1p_1^{\prime \prime })^2+ik_1q_1^{\prime \prime }\frac{1}{2}i\lambda _1k_1]`$ $`\times \mathrm{exp}[ixk_1i\lambda _2e^{cx}k_2+ix\kappa _1]`$ $`\times \mathrm{exp}[\frac{1}{2}i\lambda _1\kappa _1i\kappa _1q_1^{}\frac{1}{2}(\kappa _1p_1^{})^2]`$ $`\times dk_2dk_1dxd\kappa _1d\lambda _1d\lambda _2.`$ (152) When normalized appropriately, this expression is evaluated as $`\mathrm{exp}[\frac{1}{2}(p_1^{\prime \prime 2}+p_2^{\prime \prime 2}+icp_1^{\prime \prime })]\mathrm{exp}[\frac{1}{2}(p_1^2+p_2^2icp_1^{})],`$ (153) which once again represents a one-dimensional Hilbert space although it has a different representation than in the case $`c=0`$. Thus we have obtained a $`c`$-dependent family of distinct but equivalent quantum representations for the same Hilbert space, reflecting the $`c`$-dependent family of equivalent classical solutions. ### 6.2 Irregular constraints In discussing constraints one often pays considerable attention to the regularity of the expressions involved. Consider, once again, the simple example of a single constraint $`p=0`$ as illustrated by the classical action $`I=(p\dot{q}\lambda p)𝑑t.`$ (154) The equations of motion read $`\dot{q}=\lambda `$, $`\dot{p}=0`$, and $`p=0`$. On the other hand, one may ask about imposing the constraint $`p^3=0`$ or possibly $`p^{1/3}=0`$, etc., instead of $`p=0`$. Let us incorporate several such examples by studying the classical action $`(p\dot{q}\lambda p|p|^\gamma )𝑑t,\gamma >1.`$ (155) Here the equations of motion include $`\dot{q}=\lambda (\gamma +1)|p|^\gamma `$ which, along with the constraint $`p|p|^\gamma =0`$, may cause some difficulty in seeking a classical solution of the equations of motion. When $`\gamma 0`$, such constraints are said to be irregular . It is clear from (9) that irregular constraints lead to considerable difficulty in conventional phase-space path integral approaches. Let us examine the question of irregular constraints from the point of view of a coherent state, projection operator, phase-space path integral quantization. We first observe that the operator $`P|P|^\gamma `$ is well defined by means of its spectral decomposition. Moreover, for any $`\gamma >1`$, it follows that $`{\displaystyle e^{i\xi P|P|^\gamma }\frac{\mathrm{sin}(\delta ^{\gamma +1}\xi )}{\pi \xi }𝑑\xi }`$ $`=\mathrm{EI}(\delta ^{\gamma +1}P|P|^\gamma \delta ^{\gamma +1})`$ $`=\mathrm{EI}(\delta P\delta ).`$ (156) Thus, from the operator point of view, it is possible to consider the constraint operator $`P|P|^\gamma `$ just as easily as $`P`$ itself. In particular, it follows that $`p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}={\displaystyle \mathrm{exp}[i(p\dot{q}\lambda p|p|^\gamma )𝑑t]𝒟p𝒟q𝒟C_\gamma (\lambda )},`$ (157) where we have appended $`\gamma `$ to the measure for the Lagrange multiplier $`\lambda `$ to emphasize the dependence of that measure on $`\gamma `$. The reduction of the reproducing kernel proceeds as with the cases discussed earlier, and we determine for all $`\gamma `$ that $`\underset{\delta 0}{lim}{\displaystyle \frac{\sqrt{\pi }}{(2\delta )}}p^{\prime \prime },q^{\prime \prime }|\mathrm{EI}|p^{},q^{}=e^{{\scriptscriptstyle \frac{1}{2}}\left(p^{\prime \prime 2}+p^2\right)},`$ (158) representative of a one-dimensional Hilbert space. Note that, like the classical theory, the ultimate form of the quantum theory is independent of $`\gamma `$. It is natural to ask how one is to understand this acceptable behavior for the quantum theory for irregular constraints and the difficulties they seem to present to the classical theory. Just like the classical and quantum Hamiltonians, the connection between the classical and quantum constraints is given by $`\varphi (p,q)p,q|\mathrm{\Phi }(P,Q)|p,q=0|\mathrm{\Phi }(P+p,Q+q)|0.`$ (159) With this rule we typically find that $`\varphi (p,q)\mathrm{\Phi }(p,q)`$ due to the fact that $`\mathrm{}0`$, but the difference between these expressions is generally qualitatively unimportant. In certain circumstances, however, that difference is qualitatively significant even though it is quantitatively very small. Since that difference is $`O(\mathrm{})`$, let us explicitly exhibit the appropriate $`\mathrm{}`$-dependence hereafter. First consider the case of $`\gamma =2`$. In that case $`p,q|P^3|p,q=0|(P+p)^3|0=p^3+3P^2p,`$ (160) where we have introduced the shorthand $`()0|()|0`$. Since $`P^2=\mathrm{}/2`$ it follows that for the quantum constraint $`P^3`$, the corresponding classical constraint function is given by $`p^3+(3\mathrm{}/2)p`$. For $`|p|\sqrt{\mathrm{}}`$, this constraint is adequately given by $`p^3`$. However, when $`|p|\sqrt{\mathrm{}}`$as must eventually be the case in order to actually satisfy the classical constraint—then the functional form of the constraint is effectively $`(3\mathrm{}/2)p`$. In short, if the quantum constraint operator is $`P^3`$, then the classical constraint function is in fact regular when the constraint vanishes. A similar analysis holds for a general value of $`\gamma `$. The classical constraint is given by $`\varphi _\gamma (p)=(\pi \mathrm{})^{1/2}{\displaystyle (k+p)|k+p|^\gamma e^{k^2/\mathrm{}}𝑑k}`$ $`=(\pi \mathrm{})^{1/2}{\displaystyle k|k|^\gamma e^{(kp)^2/\mathrm{}}𝑑k}.`$ (161) For $`|p|\sqrt{\mathrm{}}`$ this expression effectively yields $`\varphi _\gamma (p)p|p|^\gamma `$. On the other hand, for $`p0`$, and more especially for $`|p|\sqrt{\mathrm{}}`$, this expression shows that the constraint function vanishes linearly, specifically as $`\varphi _\gamma (p)\kappa p`$, where $`\kappa 2(\mathrm{}^\gamma /\pi )^{1/2}{\displaystyle y^2|y|^\gamma e^{y^2}𝑑y}=2(\mathrm{}^\gamma /\pi )^{1/2}\mathrm{\Gamma }((\gamma +3)/2)\mathrm{}^{\gamma /2}\kappa _o.`$ (162) A rough, but qualitatively correct expression for this behavior is given by $`\varphi _\gamma (p)\kappa _op(\mathrm{}+p^2\kappa _o^{2/\gamma })^{\gamma /2}.`$ (163) Thus, from the present point of view, irregular constraints do not arise from consistent quantum constraints; instead, irregular constraints arise as limiting expressions of consistent, regular classical constraints as $`\mathrm{}0`$. ## 7 OTHER APPLICATIONS OF THE PROJECTION OPERATOR APPROACH There have been several cases in which the projection operator has been used to study constrained systems. Shabanov as well as Govaerts and Klauder have applied the projection operator formalism to a simple $`0+1`$ model of a gauge theory. Govaerts applied the projection operator scheme to study the relativistic particle in a reparameterization invariant form. Shabanov and Klauder have studied both first-class and second-class constraint situations from the point of view of projection operator quantization. In addition, they have discussed in a general way the application of projection operator techniques to gauge theory . Fermion systems have been treated, e.g., in . Shabanov has incorporated the projection operator into his Physics Reports review of gauge theories, and developed an algorithm for how the projection operator approach may be incorporated into lattice gauge theory calculations. Shabanov has also shown how the projection operator approach may be especially useful in ensuring constraints are satisfied in an ion-surface interaction . In addition, Klauder has applied the projection operator method in a study of quantum gravity. Finally, a U(1) Chern-Simons model has been studied and solved with the projection operator method using coherent states in . Projection operators have also been used previously in the study of constrained system quantization. For example, as noted earlier, some aspects of a coherent state quantization procedure that emphasized projection operators for systems with closed first-class constraints have been presented by Shabanov . In addition, we thank M. Henneaux for his thoughtful comments as this approach was being developed, as well as for pointing out that projection operators for closed first-class constraints also appear in the text of Henneaux and Teitelboim . Please note that this very short list does not pretend to be complete regarding prior considerations of projection operator investigations in connection with constrained systems. ## Acknowledgements The present paper represents a summary of some of the author’s principal contributions to the projection operator approach for the quantization of systems with constraints for the four years from 1996 through 1999. It is a pleasure to thank my coauthors J. Govaerts and S. Shabanov who have shared in this general project. In addition to these two individuals, thanks are also extended to M. Henneaux and B. Whiting for many discussions regarding constraints and their quantization.
warning/0003/math0003210.html
ar5iv
text
# Branching rules for modular fundamental representations of symplectic groups ## 1. Introduction The article is devoted to finding branching rules for the fundamental representations of the sympplectic groups in positive characteristic. The classical branching rules are concerned with the restrictions of representations of the classical algebraic and symmetric groups in characteristic 0 to naturally embedded subgroups of smaller ranks. For a group of rank $`n`$ and its fixed irreducible representation $`\phi `$ they yield the composition factors of the restriction of $`\phi `$ to a naturally embedded subgroup of rank $`n1`$ and hence to similar subgroups of smaller ranks, at least algorithmically. These rules provide a basis for induction on rank and have found numerous applications. In positive characteristic one cannot expect to obtain complete branching rules in an explicit form in a near future since this problem is closely connected with that of finding the dimensions of arbitrary irreducible representations and the composition factors of the Weyl modules. So it is worth to investigate important particular cases where such rules can be found and to seek for asymptotic analogs of these rules. The notion of an inductive system of representations (see the definition below) introduced by Zalesskii in yields an asymptotic version of the branching rules. It proved to be useful for the study of ideals in group algebras of locally finite groups as well, see, for instance, Zalesskii’s survey . We classify the inductive systems of the fundamental representations for the infinite-dimensional symplectic group $`Sp_{\mathrm{}}(K)`$. This class of representations yields an example of representations of a simple form for which the branching rules in positive characteristic differ from the characteristic 0 case. Let $`KF`$ be fields of characteristic $`p>0`$, $`\overline{F}`$ be the algebraic closure of $`F`$, and $`^+`$ be the set of nonnegative integers. Let $`G_n=Sp_{2n}(K)`$. Denote by $`\omega _i^n`$, $`0in`$, the $`i`$th fundamental module and representation of $`G_n`$ over $`F`$ where $`\omega _0^n`$ is the trivial one. Let $`W_0^n=\omega _0^n,W_1^n,\mathrm{},W_n^n`$ be the corresponding Weyl modules. Set $`W_i^n=\omega _i^n=0`$ for $`i<0`$ and for $`i>n`$. The labeling of the fundamental modules is standard, the fundamental and the Weyl modules for $`G_n`$ are the $`F`$-modules affording the restrictions to $`G_n`$ of the relevant representations of the group $`Sp_{2n}(\overline{F})`$ (it is well known that these restrictions can be realized over $`F`$). For an integer $`z>0`$ we denote by $`lp(z)`$ the maximal $`i`$ such that $`p^iz`$. We have $`lp(z)=0`$ if $`pz`$. Let $`M`$ be a $`G_n`$-module. The restriction of $`M`$ to $`G_{n1}`$ is denoted by $`MG_{n1}`$. We shall write $$MN_1+\mathrm{}+N_q$$ if there is a series $`0=M_0M_1\mathrm{}M_q=M`$ of submodules of $`M`$ and a permutation $`\sigma `$ such that $`N_{\sigma (i)}M_i/M_{i1}`$ for all $`i=1,\mathrm{},q`$. Moreover, if in addition $`M_i/M_{i1}`$ coincides with the socle of $`M/M_{i1}`$ for $`i=1,\mathrm{},q`$, then the sequence $$N_{\sigma (1)}_sN_{\sigma (2)}_s\mathrm{}_sN_{\sigma (q)}$$ is called the socle series of $`M`$. Theorem 1.1 below describes the branching rules for the fundamental $`G_n`$-modules and the submodule structure of the restrictions of these modules to $`G_{n1}`$. ###### Theorem 1.1. Let $`n2`$ and $`0in`$. Set $`d=lp(ni+1)`$; $`\epsilon =0`$ if $`ni+1p^d(modp^{d+1})`$ and $`\epsilon =1`$ otherwise. Then (i) $`\omega _i^nG_{n1}\omega _i^{n1}+2\omega _{i1}^{n1}+\left(_{t=0}^{d1}2\omega _{i2p^t}^{n1}\right)+\epsilon \omega _{i2p^d}^{n1}`$ (the sum in the brackets is zero whenever $`d=0`$); (ii) $`\omega _i^nG_{n1}=\omega _{i1}^{n1}\omega _{i1}^{n1}D`$ and the series $$\omega _{i2}^{n1}_s\omega _{i2p}^{n1}_s\mathrm{}_s\omega _{i2p^{d1}}^{n1}_s\omega _s\omega _{i2p^{d1}}^{n1}_s\mathrm{}_s\omega _{i2p}^{n1}_s\omega _{i2}^{n1}$$ with $`\omega =\omega _i^{n1}\epsilon \omega _{i2p^d}^{n1}`$ and $`\omega _j^{n1}`$ omitted for $`j<0`$ is the socle series of $`D`$. In particular, $`D=\omega `$ if $`i=0,1`$ or $`pn+1i`$. ###### Corollary 1.2. For $`n2`$ the restriction $`\omega _i^nG_{n1}`$ is completely reducible if and only if $`i=0,1`$ or $`pn+1i`$. The proof of Theorem 1.1 is based on the description of the composition factors of the fundamental Weyl modules (Premet and Suprunenko \[6, Theorem 2\] for $`p>2`$ and independently Adamovich \[1, Theorem 2 and its Corollary 1\] for arbitrary $`p`$) and Adamovich’s results on the submodule structure of these Weyl modules. In Section 2 these results are refined (Theorem 2.13). In particular, a new irreducibility criterion for the fundamental Weyl modules is obtained (Corollary 2.14) and it is proved that their socles are always simple (Corollary 2.15). For $`np+2in`$ Gow has given an explicit construction of the modules $`\omega _i^n`$ and has described the submodule structure of the restrictions $`\omega _i^nG_1\times G_{n1}`$ (the natural embedding) (\[5, Theorem 2.2\]). This implies our Theorem 1.1 for these modules. In a certain explicitly determined operator $`\delta `$ on the exterior algebra $`V`$ of the natural $`G_n`$-module $`V`$ is considered and it is proved that for $`np+2in`$ the module $`\omega _i^n`$ can be realized as the quotient $`\mathrm{ker}\delta ^iV/\delta ^{p1}(^{i+2p2}V)`$ (\[5, Corollary 2.4\]). This nice construction gives a realization for an important class of modules without complicated representation-theoretic machinery. However, it cannot be extended to other fundamental modules since according to \[5, Theorem 4.2\], the quotient above is zero for $`i<np+2`$. In Sheth has found the branching rules for modular representations of symmetric groups corresponding to two part partitions. The composition factors occurring in the relevant restrictions are similar to those of the module $`D`$ in Theorem 1.1(ii). We conjecture that the submodule structure of these restrictions is also similar to that of $`D`$. The authors plan to consider this question as well as the similar one for representations of special linear groups with highest weights $`\omega _i+\omega _j`$ in a subsequent paper. In Section 4 Theorem 1.1 is applied to classify the inductive systems of fundamental $`F`$-representations for $`Sp_{\mathrm{}}(K)`$. Let $$H_1H_2\mathrm{}H_n\mathrm{}$$ (1) be a sequence of groups, and $`\mathrm{\Psi }_n`$, $`n=1,2,\mathrm{}`$, be a nonempty finite set of (inequivalent) irreducible representations of $`H_n`$ over a fixed field. The system $`\mathrm{\Psi }=\{\mathrm{\Psi }_nn=1,2,\mathrm{}\}`$ is called an inductive system (of representations) for the group $`H=_{n=1}^{\mathrm{}}H_n`$ if each $`\mathrm{\Psi }_n`$ coincides with the union of the sets of composition factors (up to equivalence) of the restrictions $`\pi H_n`$ where $`\pi `$ runs over $`\mathrm{\Psi }_{n+1}`$. In this article (1) is the sequence of the naturally embedded groups $`G_n=Sp_{2n}(K)`$, so $`_{n=1}^{\mathrm{}}G_n=Sp_{\mathrm{}}(K)`$. Set $$\begin{array}{cccccc}\hfill _n& =& \{\omega _i^n0in\},\hfill & \hfill & =& \{_nn=1,2,\mathrm{}\};\hfill \\ \hfill _n^s& =& \{\omega _i^n0is\},\hfill & \hfill ^s& =& \{_n^sn=1,2,\mathrm{}\};\hfill \\ \hfill _n^u& =& \{\omega _i^nn+1uin\},\hfill & \hfill ^u& =& \{_n^un=1,2,\mathrm{}\}.\hfill \end{array}$$ ###### Theorem 1.3. The inductive systems of fundamental representations over $`F`$ for $`Sp_{\mathrm{}}(K)`$ are exhausted by the systems $``$, $`^s`$, $`^{p^t1}`$, and $`^s^{p^t1}`$ ($`s0`$, $`t1`$). It is clear that $`^0`$ (which consists of the trivial representations) and $`^{p1}`$ are minimal inductive systems. However, the question on the minimal inductive systems for $`Sp_{\mathrm{}}(K)`$ is far from solution. For $`p>2`$ Zalesskii and Suprunenko have described the inductive system $`\mathrm{\Phi }=\{\mathrm{\Phi }_nn=1,2,\mathrm{}\}`$ where for each $`n`$ the set $`\mathrm{\Phi }_n`$ consists of two irreducible representations with highest weights $`\omega _{n1}+\frac{1}{2}(p3)\omega _n`$ and $`\frac{1}{2}(p1)\omega _n`$. The system $`\mathrm{\Phi }`$ coincides with $`^2`$ for $`p=3`$ and yields another example of a minimal inductive system for $`p>3`$. For other classical groups the questions investigated in this paper present no problems since the situation is the same as in characteristic 0. The authors have found the minimal and the minimal nontrivial inductive systems for the group $`SL_{\mathrm{}}(\overline{K})`$. For this group the system consisting of the trivial representation is the only minimal inductive system, and the minimal nontrivial ones are exhausted by the systems $`𝐋^j=\{L_n^jn=1,2,\mathrm{}\}`$ and $`𝐑^j=\{R_n^jn=1,2,\mathrm{}\}`$ where $`L_n^j`$ consists of two irreducible representations of $`SL_{n+1}(\overline{K})`$ with highest weights $`0`$ and $`p^j\omega _1`$ and $`R_n^j`$ of those with highest weights $`0`$ and $`p^j\omega _n`$. The picture is similar for the groups $`SL_{\mathrm{}}`$ and $`SU_{\mathrm{}}`$ over locally finite fields. Until Proposition 4.2 we assume that $`K=F=\overline{F}`$. At the end Proposition 4.2 transfers the results to arbitrary fields. ## 2. The structure of the fundamental Weyl modules In this section we refine the results of , , and on the structure of the fundamental Weyl modules for $`G_n`$. Throughout the paper we set $`\pi _i^n=\omega _{n+1i}^n`$ and $`V_i^n=W_{n+1i}^n`$. We denote by $`[a,b]`$ the set of all $`j^+`$ with $`ajb`$. For an integer $`k^+`$ write its $`p`$-adic expansion $`k=k_0+k_1p+\mathrm{}+k_sp^s`$ with $`0k_i<p`$ and set $`k_i=0`$ for all such $`i^+`$ that $`p^i>k`$. We shall write $`k=(k_0,k_1,\mathrm{},k_s)`$. We say that an integer $`m`$ contains $`k`$ to base $`p`$ and write $`k_pm`$ if and only if for each $`i`$ either $`k_i=0`$, or $`k_i=m_i`$. Set $`d_k^m=1`$ if $`k_pm`$, and $`d_k^m=0`$ otherwise. ###### Theorem 2.1. \[6, Theorem 2\] Let $`p>2`$. Then $`W_i^n_{k=0}^{\mathrm{}}d_k^{n+1i+2k}\omega _{i2k}^n`$. We need some more notation to state Adamovich’s results. For $`\lambda ^+`$ define maps $`s_\lambda ^{}:^+^+`$ and $`s_\lambda :^+^+`$ setting $`s_\lambda ^{}(l)=l+2k^{}`$ where $`l+1=a^{}p^\lambda k^{}`$, $`a^{}^+`$, $`0k^{}<p^\lambda `$; $`s_\lambda (l)=l+2k`$ where $`l=ap^\lambda k`$, $`a^+`$, $`0k<p^\lambda `$. We say that the reflection $`s_\lambda ^{}`$ or $`s_\lambda `$ is $`l`$-admissible if $`k^{}0`$ and $`pa^{}`$ or $`k0`$ and $`pa`$, respectively. We denote by $`S(l)`$ the set of all $`m>l`$ that can be written in the form $`m=s_{\lambda _u}\mathrm{}s_{\lambda _1}(l)`$ where $`\lambda _u<\mathrm{}<\lambda _1`$ and for each $`i=0,1,\mathrm{},u1`$ the reflection $`s_{\lambda _{i+1}}`$ is $`s_{\lambda _i}\mathrm{}s_{\lambda _1}(l)`$-admissible. Similarly we define $`S^{}(l)`$ (writing $`s_{\lambda _i}^{}`$ instead of $`s_{\lambda _i}`$). ###### Theorem 2.2. Let $`0ln`$. Then $`V_{l+1}^n\pi _{l+1}^n+_{mS^{}(l)}\pi _{m+1}^n`$. As $`s_\lambda ^{}(x1)=s_\lambda (x)1`$, the following theorem yields an equivalent statement. ###### Theorem 2.3. Let $`1ln+1`$. Then $`V_l^n\pi _l^n+_{mS(l)}\pi _m^n`$. Let us rewrite Theorem 2.1 in terms of $`\pi _m^n`$ and $`V_l^n`$ (without restrictions on $`p`$). ###### Theorem 2.4. Let $`1ln+1`$. Then $`V_l^n_{k=0}^{\mathrm{}}d_k^{l+2k}\pi _{l+2k}^n`$. Now our goal is to show that Theorems 2.3 and 2.4 are equivalent, so Theorem 2.4 (and 2.1) holds for $`p=2`$. For this purpose we prove some technical facts on the triples $`k,l,m`$ with $`k_pm=l+2k`$ and admissible reflections. Until the end of the section $`l1`$. For each $`mS(l)`$ the tuple $`(\lambda _1;\mathrm{};\lambda _u)`$ is uniquely determined (see the comments before the Theorem in ). If $`u`$ is odd for some $`m`$, set $`\lambda _{u+1}=lp(m)`$. Then $`s_{\lambda _{u+1}}(m)=m`$ and $`\lambda _{u+1}<\lambda _u`$. Now for every $`mS(l)`$ we have a uniquely determined sequence of reflections $`s_{\lambda _1},\mathrm{},s_{\lambda _{2t}}`$. Such sequences will be called $`l`$-admissible. For an integer $`0ap1`$ set $`\overline{a}=p1a`$. The following lemma is straightforward. ###### Lemma 2.5. Set $`q=lp(l)`$. The reflection $`s_\lambda `$ is $`l`$-admissible if and only if $`\lambda >q`$ and $`l_\lambda p1`$. In that case $`s_\lambda (l)=p^q(\overline{l}_q+1,\overline{l}_{q+1},\mathrm{}\overline{l}_{\lambda 1},l_\lambda +1,l_{\lambda +1},\mathrm{}).`$ Two consequent applications of Lemma 2.5 yield ###### Proposition 2.6. Let $`lp(l)\mu <\lambda `$, $`m=s_\mu s_\lambda (l)`$, and $`k=(ml)/2`$. The pair $`s_\lambda ,s_\mu `$ is $`l`$-admissible if and only if $`l_\lambda p1`$ and $`l_\mu 0`$. In that case $$\begin{array}{cccccccccccc}\hfill m& =& (l_0,\hfill & \hfill \mathrm{},& \hfill l_{\mu 1},& \hfill \overline{l}_\mu +1,& \overline{l}_{\mu +1},\hfill & \mathrm{},\hfill & \overline{l}_{\lambda 1},\hfill & l_\lambda +1,\hfill & l_{\lambda +1},\hfill & \mathrm{}),\hfill \\ \hfill k& =& (0,\hfill & \hfill \mathrm{},& \hfill 0,& \hfill \overline{l}_\mu +1,& \overline{l}_{\mu +1},\hfill & \mathrm{},\hfill & \overline{l}_{\lambda 1}).\hfill & & & \end{array}$$ In particular, $`k_pm=l+2k`$. We call a tuple $`\sigma =(\lambda _1;\mathrm{};\lambda _{2t})`$ $`l`$-admissible if $`\lambda _i^+`$, $`\lambda _1>\mathrm{}>\lambda _{2t}`$, $`l_{\lambda _{2j1}}p1`$ and $`l_{\lambda _{2j}}0`$ for $`j=1,\mathrm{},t`$. For an $`l`$-admissible $`\sigma `$ set $$𝔔_l(\sigma )=\underset{j=1}{\overset{t}{}}[\lambda _{2j},\lambda _{2j1}1],$$ $`\delta _{i\sigma }=1`$ if $`i=\lambda _j`$ for some $`j`$ and $`\delta _{i\sigma }=0`$ otherwise. Define $`l^\sigma ^+`$ putting $$l_i^\sigma =\{\begin{array}{cc}\overline{l}_i+\delta _{i\sigma },\hfill & i𝔔_l(\sigma )\hfill \\ l_i+\delta _{i\sigma },\hfill & i𝔔_l(\sigma ).\hfill \end{array}$$ Proposition 2.6 yields the following corollary. ###### Corollary 2.7. A sequence $`s_{\lambda _1},\mathrm{},s_{\lambda _{2t}}`$ is $`l`$-admissible if and only if the tuple $`\sigma =(\lambda _1;\mathrm{};\lambda _{2t})`$ is $`l`$-admissible. In that case $`s_{\lambda _{2t}}\mathrm{}s_{\lambda _1}(l)=l^\sigma `$. ###### Proposition 2.8. An integer $`mS(l)`$ if and only if $`ml=2k>0`$ and $`k_pm`$. ###### Proof. Let $`\sigma =(\lambda _1;\mathrm{};\lambda _{2t})`$ be an $`l`$-admissible tuple and $`m=s_{\lambda _{2t}}\mathrm{}s_{\lambda _1}(l)`$. Set $`m^0=l`$, $`m^j=s_{\lambda _{2j}}\mathrm{}s_{\lambda _1}(l)`$, and $`k^j=(m^jm^{j1})/2`$, $`1jt`$. Using Proposition 2.6, one deduces that $`k^j_pm^j`$; $`m^j=l^{\sigma ^j}`$ with $`\sigma ^j=(\lambda _1;\mathrm{};\lambda _{2j})`$; $`k_i^j=0`$ and $`m_i^j=m_i^{j1}`$ for $`i[\lambda _{2j},\lambda _{2j1}]`$. Therefore $`k=k^1+\mathrm{}+k^t_pm`$. Assume now that $`ml=2k>0`$ and $`k_pm`$. Choose integers $`\tau _1,\tau _2,\mathrm{},\tau _{2t}`$ as follows. Set $`\tau _0=1`$. Assume that $`\tau _{2j}`$ is chosen. If there is no $`i>\tau _{2j}`$ such that $`k_i=m_i0`$, we set $`t=j`$ and stop the process. Otherwise we choose for $`\tau _{2j+1}`$ minimal $`i>\tau _{2j}`$ with $`k_i=m_i0`$. As $`m>2k`$, there exists $`f>\tau _{2j+1}`$ with $`k_fm_f`$ (observe that in this case $`k_f=0`$). We choose minimal such $`f`$ for $`\tau _{2j+2}`$. Set $`\lambda _q=\tau _{2t+1q}`$. Since $`k_pm`$, using Corollary 2.7 and analyzing the $`p`$-adic expansions of $`k`$ and $`m`$, one can conclude that the tuple $`\sigma =(\lambda _1;\mathrm{};\lambda _{2t})`$ is $`l`$-admissible and $`m=l^\sigma `$. ∎ ###### Corollary 2.9. Theorems 2.3 and 2.4 are equivalent, so Theorems 2.1 and 2.4 are valid in characteristic 2 as well. Now we rewrite Adamovich’s results on the submodule structure of the Weyl modules in our terms. We fix $`n`$ and write $`V_l`$ and $`\pi _m`$ instead of $`V_l^n`$ and $`\pi _m^n`$. For $`mS(l)`$ or $`m=l`$ we denote by $`P_l(m)`$ the smallest submodule of $`V_l`$ that has a composition factor $`\pi _m`$. Since $`V_l`$ is multiplicity-free, $`P_l(m)`$ is correctly defined and each submodule of $`V_l`$ is a sum of $`P_l(m)`$ for some $`m`$. Hence the submodule structure of $`V_l`$ is determined by the inclusion relations between the submodules $`P_l(m)`$ (see also comments at the beginning of ). We shall write $`\pi _m\pi _q`$ if $`P_l(m)P_l(q)`$. Let $`\sigma =(\lambda _1;\mathrm{};\lambda _{2t})`$ be an $`l`$-admissible tuple. For $`m=l^\sigma `$ set $$𝔔_l(m)=𝔔_l(\sigma )=\underset{j=1}{\overset{t}{}}[\lambda _{2j},\lambda _{2j1}1].$$ Note that $`𝔔_l(m)=_{j=1}^t[\tau _{2j1},\tau _{2j}1]`$ where $`\tau _i`$ are as in the proof of Proposition 2.8. For instance, for $`p=3`$, $`m`$ $`=`$ $`(0,1,\underset{¯}{2,2,0},1,0,\underset{¯}{2,1,0,0},2,1),`$ $`k`$ $`=`$ $`(0,0,\underset{¯}{2,2,0},0,0,\underset{¯}{2,1,0,0},0,0),`$ we have $`𝔔_{m2k}(m)=[2,4][7,10]`$. Put also $`𝔔_l(l)=\mathrm{}`$. ###### Theorem 2.10. For $`1m,qn+1`$ and $`m,qS(l)\{l\}`$ the module $`\pi _m\pi _q`$ (as composition factors of $`V_l`$) if and only if $`𝔔_l(q)𝔔_l(m)`$. ###### Remark 2.11. Actually the sets $`𝔓_l(m)`$ which are considered in differ slightly from $`𝔔_l(m)`$. For $`mS(l)`$ one has $`𝔓_l(m)=_{j=1}^t[\mu _{2j}+1,\mu _{2j1}]`$ where $`\mu _s=\lambda _s`$ for $`s<2t`$, $`\mu _{2t}=\lambda _{2t}`$ if $`ms_{\lambda _{2t1}}\mathrm{}s_{\lambda _1}(l)`$, and $`\mu _{2t}=0`$ otherwise. However, Lemma 2.5 enables one to deduce that $`𝔓_l(m)𝔓_l(q)`$ if and only if $`𝔔_l(m)𝔔_l(q)`$. The crucial point is that $`lp(l)=lp(m)`$ for $`mS(l)`$. For $`l`$-admissible tuples $`\sigma =(\lambda _1;\mathrm{};\lambda _{2t})`$ and $`\sigma ^{}=(\lambda _1^{};\mathrm{};\lambda _{2s}^{})`$ we say that $`\sigma \sigma ^{}`$ if there exists $`f2t,2s`$ such that $`\lambda _i=\lambda _i^{}`$ for $`1if`$ and either $`f=2t`$, or $`f<2t,2s`$ and $`\lambda _{f+1}<\lambda _{f+1}^{}`$. It is convenient to assume that the empty tuple $`\mathrm{}`$ is $`l`$-admissible, $`\mathrm{}\sigma `$ for all $`\sigma `$, $`l^{\mathrm{}}=l`$, and $`𝔔_l(\mathrm{})=\mathrm{}`$. The following is obvious. ###### Lemma 2.12. Let $`\sigma `$ and $`\sigma ^{}`$ be $`l`$-admissible tuples. Then $`l^\sigma l^\sigma ^{}`$ if and only if $`\sigma \sigma ^{}`$. Set $`n^{}=n+1`$. Construct an $`l`$-admissible tuple $`\sigma ^{\mathrm{max}}=(\mu _1;\mathrm{};\mu _{2t})`$ as follows. Put $`\mu _0=+\mathrm{}`$. Assume that $`\mu _{2j}`$ is chosen. Set $`\mu =\mu _{2j}1`$. If there is no $`l`$-admissible tuple $`(\alpha ;\beta )`$ such that $`\mu \alpha >\beta `$ and $$(l_0,\mathrm{},l_\mu )^{(\alpha ;\beta )}=(l_0,\mathrm{},\overline{l}_\beta +1,\overline{l}_{\beta +1},\mathrm{},\overline{l}_{\alpha 1},l_\alpha +1,\mathrm{},l_\mu )(n_0^{},\mathrm{},n_\mu ^{}),$$ we stop the process and set $`t=j`$ ($`\sigma ^{\mathrm{max}}=\mathrm{}`$ if $`t=0`$). Otherwise we choose maximal such pair $`(\alpha ;\beta )`$ (with respect to $``$); set $`\mu _{2j+1}=\alpha `$ and $`\mu _{2j+2}=\beta `$; and if $$(\overline{l}_\beta +1,\overline{l}_{\beta +1},\mathrm{},\overline{l}_{\alpha 1},l_\alpha +1)<(n_\beta ^{},\mathrm{},n_\alpha ^{}),$$ we stop the process and determine $`(\mu _{2j+3};\mathrm{};\mu _{2t})`$ as the maximal $`l`$-admissible tuple with $`\mu _{2j+3}<\beta `$. Obviously, $`l^{\sigma ^{\mathrm{max}}}`$ is the maximal integer $`m`$ such that $`\pi _m^n`$ is a composition factor of $`V_l^n`$. For $`l`$-admissible tuples $`\sigma `$ and $`\sigma ^{}`$ we write $`\sigma \sigma ^{}`$ if and only if $`𝔔_l(\sigma )𝔔_l(\sigma ^{})`$. Using Corollary 2.7, Theorem 2.10, and Lemma 2.12, we get our main result on the structure of fundamental Weyl modules. ###### Theorem 2.13. The map $`\sigma \pi _{l^\sigma }^n`$ is a poset isomorphism between the $`l`$-admissible tuples $`\sigma \sigma ^{\mathrm{max}}`$ and the composition factors of $`V_l^n`$ with the partial orders $``$. If $`l<n^{}`$, we denote by $`v`$ the maximal integer such that $`l_vn_v^{}`$. If $`l_v+1=n_v^{}`$, we denote by $`u`$ the maximal integer $`<v`$ such that $`\overline{l}_un_u^{}`$ setting $`u=1`$ if $`\overline{l}_i=n_i^{}`$ for all $`i<v`$. Put $`s=lp(l)`$. ###### Corollary 2.14. Let $`n^{}=n+1`$ and $`1ln^{}`$. Then $`V_l^n`$ is irreducible (i.e. $`\sigma ^{\mathrm{max}}=\mathrm{}`$) if and only if one of the following holds. (1) $`l=n^{}`$; (2) $`l<n^{}`$ and $`sv`$; (3) $`l<n^{}`$, $`s<v`$, $`l_v+1=n_v^{}`$, $`\overline{l}_sn_s^{}`$; $`l_i=p1`$ and $`n_i^{}=0`$ for $`s<i<v`$. ###### Proof. This follows from Proposition 2.6 and Corollary 2.7. ∎ ###### Corollary 2.15. Let $`n^{}=n+1`$ and $`1ln^{}`$. The socle of $`V_l^n`$ is always simple. For reducible $`V_l^n`$ it has the form $`\pi _{l^\gamma }`$ with $`\gamma =(t;s)`$ and $`t`$ as follows. (1) $`t=v`$ if $`s<v`$ and either $`l_v+1<n_v^{}`$, or $`\overline{l}_u<n_u^{}`$; (2) $`t=w`$ if $`l_v+1=n_v^{}`$; $`u=1`$ or $`\overline{l}_u>n_u^{}`$; $`s<w<v`$; $`l_wp1`$; and $`l_j=p1`$ for $`w<j<v`$. ###### Proof. Applying Results 2.6, 2.7, and 2.13, we conclude that $`\pi _{l^\gamma }`$ is a composition factor of $`V_l^n`$ and for each $`l`$-admissible tuple $`\tau \sigma ^{\mathrm{max}}`$ the set $`𝔔_l(\tau )[s,t1]`$, so $`\pi _{l^\gamma }\pi _{l^\tau }`$. ∎ ## 3. Branching rules and the submodule structure of the restrictions In this section the main results of the article are proved. We shall need the following simple lemma. ###### Lemma 3.1. Assume that $`d_k^{l+2k}=1`$ (i.e. $`k_pl+2k`$). (i) If $`p^sl+2k`$, then $`p^sk`$ and $`p^sl`$. (ii) If $`p^sl`$, then $`p^sk`$ and $`p^sl+2k`$. ###### Proof. One can assume that $`s1`$. $`(i)`$ Let $`p^sl+2k`$. Since $`k_pl+2k`$, we have $`p^sk`$. This implies that $`p^sl`$. $`(ii)`$ Let $`p^sl`$. Then $`l_0=\mathrm{}=l_{s1}=0`$. Let $`r=lp(k)`$. Assume that $`r<s`$. Since $`k_pl+2k`$, we have $`k_r=(2k)_r0`$, which is impossible. Therefore $`rs`$, so $`p^sk`$ and $`p^sl+2k`$. ∎ As in Section 2 , we shall omit the superscript $`n`$ in our notation for modules when it is known which group is considered. Replacing $`\omega _i`$ by $`\pi _{n+1i}`$ and $`W_i`$ by $`V_{n+1i}`$, one immediately concludes that Theorem 1.1(i) is equivalent to the following ###### Theorem 3.2. Let $`1in+1`$ and $`d=lp(i)`$. Then $$\pi _i^nG_{n1}\pi _{i1}+2\pi _i+\left(\underset{t=0}{\overset{d1}{}}2\pi _{i1+2p^t}\right)+\epsilon \pi _{i1+2p^d}$$ where $`\epsilon =0`$ if $`ip^d(modp^{d+1})`$ and $`\epsilon =1`$ otherwise. ###### Proof. One can rewrite the formula in Theorem 3.2 as follows. $$\pi _i^nG_{n1}\pi _{i1}+2\pi _i+\underset{t=0}{\overset{\mathrm{}}{}}b_t^i\pi _{i1+2p^t}$$ (2) where $$b_t^i=\{\begin{array}{cc}\hfill 2,& i0(modp^{t+1}),\hfill \\ \hfill 1,& iap^t(modp^{t+1})\text{ and }a0,1(modp),\hfill \\ \hfill 0,& i0(modp^t)\text{ or }ip^t(modp^{t+1}).\hfill \end{array}$$ (3) Recall that by convention $`\pi _i^n=0`$ for all $`i>n+1`$, and $`\pi _{n+1}^n`$ is the trivial one-dimensional $`G_n`$-module. So (2) holds for $`in+1`$. Assume now that $`1l<n+1`$ and (2) is valid for all $`i>l`$. We shall prove it for $`i=l`$. Then the theorem will follow by induction. It follows from \[4, Proposition 3.3.2 and Theorem 4.3.1\] that $`V_lG_{n1}`$ has a filtration by Weyl modules for $`G_{n1}`$. Then the classical branching rules for characteristic 0 and Theorem 2.4 imply $$V_l^nG_{n1}V_{l1}+2V_l+V_{l+1}2𝒱+\underset{t=0}{\overset{\mathrm{}}{}}f_t^{l+2t1}\pi _{l+2t1}$$ (4) where $`𝒱=_{k=0}^{\mathrm{}}d_k^{l+2k}\pi _{l+2k}`$, $`f_0^{l1}=d_0^{l1}`$, and $$f_t^{l+2t1}=d_t^{l+2t1}+d_{t1}^{l+2t1}\text{for }t1.$$ On the other hand, by Theorem 2.4, $`V_lG_{n1}_{k=0}^{\mathrm{}}d_k^{l+2k}(\pi _{l+2k}G_{n1}).`$ Since $`d_0^l=1`$ and the branching rules for $`\pi _i`$ with $`i>l`$ are assumed to satisfy (2), one can determine the branching of $`\pi _l`$. Therefore it suffices to check that the right part of (4) is equal to $`_{k=0}^{\mathrm{}}d_k^{l+2k}U_{l+2k}`$ where $`U_i`$ is the right part of (2). The latter sum can be rewritten as follows: $$2𝒱+\underset{k=0}{\overset{\mathrm{}}{}}d_k^{l+2k}(\pi _{l+2k1}+\underset{s=0}{\overset{\mathrm{}}{}}b_s^{l+2k}\pi _{l+2k+2p^s1})=2𝒱+\underset{t=0}{\overset{\mathrm{}}{}}e_t^{l+2t1}\pi _{l+2t1}$$ where $$e_t^{l+2t1}=d_t^{l+2t}+\underset{k,s0,k+p^s=t}{}d_k^{l+2k}b_s^{l+2k}.$$ (5) We have to show that $`e_t^{l+2t1}=f_t^{l+2t1}`$ for all $`t0`$. Note that $`f_t^{l+2t1}2`$. We proceed by steps. Step 1. At most one summand in (5) is nonzero. In particular, $`e_t^{l+2t1}2`$. Assume that $`d_k^{l+2k}b_s^{l+2k}0`$ and $`d_k^{}^{l+2k^{}}b_s^{}^{l+2k^{}}0`$ with $`t=k+p^s=k^{}+p^s^{}`$ and $`s>s^{}`$. Since $`b_s^{l+2k}0`$, we have $`p^sl+2k`$. As $`d_k^{l+2k},d_k^{}^{l+2k^{}}0`$, by Lemma 3.1, $`p^s`$ divides $`k`$, $`l`$, and $`k^{}`$. Hence $`p^sk+p^sk^{}=p^s^{}`$, which yields a contradiction. Now assume that $`d_t^{l+2t}0`$ and $`d_k^{l+2k}b_s^{l+2k}0`$ with $`k+p^s=t`$. As above, we get that $`p^s`$ divides $`k`$, $`l`$, and $`t`$. Let $`r=lp(l)`$. Then by Lemma 3.1 (ii), $`p^rk`$ and $`p^rt`$. Since $`k+p^s=t`$, we have $`r=s`$, so $`l_s0`$. Consider the following cases. Case 1. $`k0`$ and $`t0(modp^{s+1})`$. Then $`l+2tt`$ and $`l+2kk(modp^{s+1})`$, so $`p^s=tk2(tk)(modp^{s+1})`$, which is impossible. Case 2. $`k0(modp^{s+1})`$. Then $`t_s=1`$. Since $`l+2tt(modp^{s+1})`$, we have $`l_s=p1`$, so $`(l+2k)_s=p1`$. This implies that $`b_s^{l+2k}=0`$ and yields a contradiction. Case 3. $`t0(modp^{s+1})`$. Then $`k_s=p1`$. Therefore $`(l+2k)_s=k_s=p1`$, so as above, $`b_s^{l+2k}=0`$. Step 2. $`e_t^{l+2t1}=2`$ if and only if $`f_t^{l+2t1}=2`$ (equivalently, $`d_t^{l+2t1}=d_{t1}^{l+2t1}=1`$). Assume that $`e_t^{l+2t1}=2`$. By Step 1, this is equivalent to the following: there exist $`k,s0`$ with $`k+p^s=t`$ such that $`b_s^{l+2k}=2`$ and $`d_k^{l+2k}=1`$. Hence $`p^{s+1}l+2k`$ and $`k_pl+2k`$. By Lemma 3.1 (i), $`p^{s+1}`$ divides $`l`$ and $`k`$. Note that $`l+2t1=l+2k+p^s+(p^s1)`$. Therefore $`t=k+p^s_pl+2t1`$ and $`t1=k+(p^s1)_pl+2t1`$, so $`d_t^{l+2t1}=d_{t1}^{l+2t1}=1`$, as required. Assume now that $`d_t^{l+2t1}=d_{t1}^{l+2t1}=1`$. Let $`s=lp(t)`$. Then $`t_s0`$ and $`(t1)_s=t_s1`$. Since both $`t`$ and $`t1`$ are contained in $`l+2t1`$, we have $`t_s=1`$. Moreover, we have $`(l1)_0=\mathrm{}=(l1)_{s1}=(t1)_0=\mathrm{}=(t1)_{s1}=p1`$. Since $`(l+2t1)_s=t_s=1`$, we get $`(l1)_s=p1`$. Hence $`p^{s+1}l`$. Set $`k=tp^s`$. Then $`p^{s+1}k`$, so $`p^{s+1}l+2k`$ and $`b_s^{l+2k}=2`$. It remains to observe that $`k=tp^s_p(l+2t1)p^s(p^s1)`$, so $`d_k^{l+2k}=1`$ and $`e_t^{l+2t1}=d_k^{l+2k}b_s^{l+2k}=2`$. Step 3. If $`e_t^{l+2t1}0`$, then $`f_t^{l+2t1}0`$, i.e. $`t_pl+2t1`$ or $`t1_pl+2t1`$. Let $`r=lp(l)`$. First assume that $`d_t^{l+2t}=1`$ (see (5)), i.e. $`t_pl+2t`$. Then by Lemma 3.1 (ii), $`p^rt`$. One easily checks that if $`p^{r+1}t`$, then $`t1_pl+2t1`$, and if $`p^{r+1}t`$, then $`t_pl+2t1`$, as required. Assume now that there exist $`k,s0`$ with $`k+p^s=t`$ such that $`d_k^{l+2k}=1`$ and $`b_s^{l+2k}0`$. By Lemma 3.1, $`p^rk`$ and $`lp(l+2k)=r`$. If $`b_s^{l+2k}=2`$, then by Step 2, $`f_t^{l+2t1}=20`$. Hence we can assume that $`b_s^{l+2k}=1`$. By (3), $`p^sl+2k`$ and $`(l+2k)_s0,p1`$, so $`r=s`$. Assume that $`k_s=0`$. Then $`(t1)_s=0`$ and $`t1=k+p^s1_pl+2k+p^s+(p^s1)=l+2t1`$. If $`k_s0`$, we have $`k_s=(l+2k)_sp1`$, so $`t=k+p^s_pl+2k+p^s+(p^s1)=l+2t1`$, as required. Step 4. If $`f_t^{l+2t1}0`$, then $`e_t^{l+2t1}0`$. If $`t=0`$, then $`e_0^{l1}=1`$, so assume that $`t1`$. We have either $`t_pl+2t1`$, or $`t1_pl+2t1`$. One needs to show that either $`t_pl+2t`$ (i.e. $`d_t^{l+2t}=1`$), or there exist $`k,s0`$ with $`k+p^s=t`$ such that $`k_pl+2k`$, $`p^sl+2k`$, and $`(l+2k)_sp1`$ (i.e. $`d_k^{l+2k}=1`$ and $`b_s^{l+2k}0`$). First assume that $`t_pl+2t1`$. Let $`s=lp(t)`$. We have $`(l+2t1)_s=t_s0`$. If $`p^sl+2t`$, then $`k=tp^s_pl+2t1p^s(p^s1)=l+2k`$, $`p^sl+2t2p^s=l+2k`$, and $`(l+2k)_s=(l+2t1)_s1p1`$, as required. If $`p^sl+2t`$, one gets $`t_p(l+2t1)+1=l+2t`$, as desired. Assume now that $`t1_pl+2t1`$. Consider the following cases. Case 1. $`(t1)_0=0`$. If $`(l+2t1)_0=0`$, then $`t_pl+2t`$. Assume that $`(l+2t1)_00`$. Set $`s=0`$, $`k=t1`$. Then $`k=t1_p(l+2t1)1=l+2k`$ and $`(l+2k)_0=(l+2t1)_01p1`$, as required. Case 2. $`(t1)_00,p1`$. Then $`(l+2t1)_0=(t1)_0p1`$, so $`t_pl+2t`$. Case 3. $`pt`$. Let $`s=lp(t)`$. Since $`t1_pl+2t1`$, we have $`p^sl+2t`$, so $`p^sl`$. As $`p^{s+1}t`$, the integer $`(t1)_sp1`$. If $`(t1)_s0`$, then $`(l+2t1)_s=(t1)_sp1`$, so $`t_pl+2t`$. Assume now that $`(t1)_s=0`$. This implies $`t_s=1`$. If $`(l+2t1)_s=0`$, then $`t_pl+2t`$. Therefore one can suppose that $`(l+2t1)_s1`$. Then $`k=tp^s_p(l+2t1)p^s(p^s1)=l+2k`$, $`p^sl+2t2p^s=l+2k`$, and $`(l+2k)_s=(l+2t1)_s1p1`$, as required. ∎ Now we investigate the submodule structure of the restriction $`\pi _i^nG_{n1}`$. Let $`n>1`$ and $`1in`$. As $`\pi _i^n`$ is the top composition factor of $`V_i^n`$, it follows from (4) that $`\pi _i^nG_{n1}`$ is a quotient of the $`G_{n1}`$-module $`V_i^nG_{n1}V_{i1}+2V_i+V_{i+1}`$. Applying Smith’s theorem both to $`V_i`$ and $`\pi _i`$, we conclude that $`V_i^nG_{n1}=V_iV_iV`$ where $`VV_{i1}+V_{i+1}`$, and $`\pi _i^nG_{n1}=\pi _i\pi _iD`$ where $`D`$ is a quotient of $`V`$. Now Theorem 1.1(ii) and Corollary 1.2 follow immediately from ###### Theorem 3.3. Let $`d=lpi`$. Set $`\epsilon =0`$ if $`ip^d(modp^{d+1})`$ and $`\epsilon =1`$ otherwise; $`j_q=i1+2p^q`$. Choose minimal $`t^+`$ such that $`j_t>n`$. Put $`d^{}=\mathrm{min}\{d,t\}`$. Then $$\pi _{j_0}_s\pi _{j_1}_s\mathrm{}_s\pi _{j_{d^{}1}}_s\pi _{i1}\epsilon \pi _{j_d}_s\pi _{j_{d^{}1}}_s\mathrm{}_s\pi _{j_1}_s\pi _{j_0}$$ is the socle series of $`D`$. In particular, $`D=\pi _{i1}\epsilon \pi _{j_d}`$ if $`d^{}=0`$. ###### Proof. By Theorem 3.2, $`D\pi _{i1}+2\pi _{j_0}+\mathrm{}+2\pi _{j_{d^{}1}}+\epsilon \pi _{j_d}.`$ It follows from Theorem 2.4 that the factors $`\pi _{i1},\pi _{j_0},\mathrm{},\pi _{j_{d^{}1}}`$ come from $`V_{i1}`$ and the factors $`\pi _{j_0},\mathrm{},\pi _{j_{d^{}1}}`$, and $`\epsilon \pi _{j_d}`$ if nonzero come from $`V_{i+1}`$. Note that $$j_k=i1+2p^k=i+p^k+(p^k1)=i+(\underset{k}{\underset{}{p1,\mathrm{},p1}},1),$$ so $`𝔔^{i1}(j_k)=[k,d1]`$ for all $`0kd1`$. Therefore by Theorem 2.10, $$\pi _{j_0}\pi _{j_1}\mathrm{}\pi _{j_{d^{}1}}\pi _{i1}\text{in }V_{i1}.$$ (6) Similarly, we get $`𝔔^{i+1}(j_k)=[0,k1]`$ for $`1k<d`$ (and for $`k=d`$ if $`\epsilon 0`$ and $`d>0`$). Hence $$\epsilon \pi _{j_d}\pi _{j_{d^{}1}}\mathrm{}\pi _{j_1}\pi _{j_0}\text{in }V_{i+1}.$$ (7) (Here the symbol $``$ is extended to the zero module in the natural way.) Since $`\pi _i^n`$ is selfdual, $`D`$ is selfdual also. Let $`D_1_s\mathrm{}_sD_m`$ be the socle series of $`D`$. Recall that $`D`$ has a filtration by quotients of $`V_{i1}`$ and $`V_{i+1}`$. As the factor $`\pi _{i1}`$ has multiplicity 1 and $`D`$ is selfdual, (6) implies that $`\pi _{i1}`$ is a factor of $`D_q`$ with $`d^{}+1qmd^{}`$, so $`m2d^{}+1`$. If $`\epsilon \pi _{j_d}=0`$, then $`m=2d^{}+1`$ is the composition length of $`D`$ and the theorem follows from (6) and (7). Assume that $`\epsilon \pi _{j_d}0`$. As above, by the selfduality of $`D`$ and (7), $`\pi _{j_d}`$ is a factor of $`D_q^{}`$ with $`d^{}+1q^{}md^{}`$. Assume that $`q^{}q`$. Then $`m=2d^{}+2`$, so $`D`$ is uniserial, which contradicts the selfduality of $`D`$. Hence $`q^{}=q`$ and the theorem follows from (6) and (7). ∎ ###### Remark 3.4. Obviously, if $`\epsilon \pi _{j_d}=0`$ (i.e. $`d^{}<d`$ or $`ip^d(modp^{d+1})`$), then the module $`D`$ is uniserial, so has exactly $`2d^{}+2`$ different submodules. Since $`D`$ is selfdual, one can easily observe that $`D`$ has exactly $`2d+4`$ different submodules in the case where $`\epsilon \pi _{j_d}0`$ (i.e. $`d=d^{}`$ and $`ip^d(modp^{d+1})`$). ## 4. Inductive systems and the transfer to an arbitrary field In this section the inductive systems of fundamental representations for $`Sp_{\mathrm{}}(K)`$ are classified and the results of the paper are transferred to an arbitrary field. ###### Proposition 4.1. (i) Let $`\mathrm{\Psi }`$ be an inductive system and $`^{p^{t1}}\mathrm{\Psi }`$. Then $`^{p^t1}\mathrm{\Psi }`$. (ii) $`^k`$ is an inductive system if and only if $`k=p^t1`$, $`t1`$. ###### Proof. (i) Assume that $`^{p^t1}\mathrm{\Psi }`$. Choose maximal $`l`$ such that $`^l\mathrm{\Psi }`$. Then $`p^{t1}l<p^t1`$. Take minimal $`s`$ such that $`l_sp1`$ and set $`i=p^s(l_s,l_{s+1},\mathrm{})`$. Then $`i>0`$ and $`^i\mathrm{\Psi }`$. One has $`lp(i)s`$. Moreover, if $`lp(i)=s`$, then $`ip^s(modp^{s+1})`$. Therefore Theorem 3.2 implies that $`\pi _{i1+2p^s}^{n1}`$ is a composition factor of $`\pi _i^nG_{n1}`$ for $`ni1+2p^s`$, and $`^{i1+2p^s}\mathrm{\Psi }`$. As $`i1+2p^s>l`$, we get a contradiction. (ii) In view of (i), it suffices to verify that $`^{p^t1}`$ is an inductive system. By Theorem 3.2, we need only to check that if $`i<p^t1`$ and $`s=lp(i)`$, then $`i1+2p^sp^t1`$ if $`i_sp1`$, and $`i1+2p^{s1}p^t1`$ if $`i_s=p1`$ and $`s>0`$. But this is clear since $`i(p2)p^s+(p1)p^{s+1}+\mathrm{}+(p1)p^{t1}`$ in the first case and $`i(p1)p^s+(p1)p^{s+1}+\mathrm{}+(p1)p^{t1}`$ in the second one. ∎ ###### Proof of Theorem 1.3.. Theorem 1.1(i) and Proposition 4.1 yield that $``$, $`^s`$, $`^{p^t1}`$, and $`^s^{p^t1}`$, $`s0`$, $`t1`$, are inductive systems for $`Sp_{\mathrm{}}(K)`$. Let $`\mathrm{\Psi }=\{\mathrm{\Psi }_i,i=1,2,\mathrm{}\}`$ be an inductive system of fundamental representations. It is clear that either for every $`s,u^+`$ there exist $`n`$ and $`l`$ such that $`\omega _l^n\mathrm{\Psi }_n`$, $`l>s`$, and $`n+1l>u`$, or $`\mathrm{\Psi }^s^u`$ for some $`s`$ and $`u`$. In the first case we claim that $`\mathrm{\Psi }=`$. Indeed, fix $`m`$ and $`l`$, $`0lm`$. Then one can choose $`k`$ and $`n`$ such that $`\omega _k^n\mathrm{\Psi }_n`$, $`kl`$, and $`nkml`$. Since $`\mathrm{\Psi }`$ is an inductive system, Theorem 1.1(i) implies that $`\omega _l^{n+lk}\mathrm{\Psi }_{n+lk}`$ and $`\omega _l^m\mathrm{\Psi }_m`$. Hence $`\mathrm{\Psi }=`$. Next, suppose that $`\mathrm{\Psi }^s^u`$. Choose minimal $`s`$ and $`u`$ with this property assuming that $`s=1`$ if $`\mathrm{\Psi }^u`$ and $`u=0`$ if $`\mathrm{\Psi }^s`$. (Observe that for all $`s`$ and $`u`$, $`(^s^u)_n=\mathrm{}`$ for $`n`$ large enough.) We shall prove that $`\mathrm{\Psi }=^s^u`$ and $`u=p^t1`$ with $`t^+`$ (in particular, $`\mathrm{\Psi }=^u`$ for $`s=1`$ and $`\mathrm{\Psi }=^s`$ for $`u=0`$). First let $`u>0`$. We claim that $`^u\mathrm{\Psi }`$ and $`u=p^t1`$. As $`\mathrm{\Psi }^s`$ and $`\mathrm{\Psi }`$ and $`^s`$ are inductive systems, $`\mathrm{\Psi }_n_n^u\mathrm{}`$ for infinitely many integers $`n`$. So there exists $`vu`$ such that $`\pi _v^n\mathrm{\Psi }_n`$ for infinitely many $`n`$. Choose maximal such $`v`$. Theorem 3.2 yields that $`\pi _v^n\mathrm{\Psi }_n`$ for all $`nv1`$ and $`^v\mathrm{\Psi }`$. Now Proposition 4.1 and the choice of $`v`$ imply that $`v=p^t1`$ and $`^v`$ is an inductive system. It remains to show that $`v=u`$. Suppose this is not the case. As $`\mathrm{\Psi }^s^v`$, there exist $`l`$ and $`t`$ such that $`v<lu`$, $`t>s+l1`$, and $`\pi _l^t\mathrm{\Psi }_t`$. Since $`\mathrm{\Psi }`$, $`^s`$, and $`^v`$ are inductive systems, this implies that for every $`k>t`$ there exists $`\pi _{m_k}^k\mathrm{\Psi }_k`$ with $`v<m_ku`$ which contradicts the choice of $`v`$. Hence $`v=u=p^t1`$ and $`^u\mathrm{\Psi }`$. Now we show that $`^s\mathrm{\Psi }`$ if $`s0`$. As $`\mathrm{\Psi }^{s1}^u`$, for some $`n>s+u1`$ we have $`\omega _s^n\mathrm{\Psi }_n`$. Since $`\mathrm{\Psi }`$, $`^u`$, and $`^{s1}`$ for $`s1`$ are inductive systems, this forces $`\omega _s^n\mathrm{\Psi }_n`$ for all $`ns`$. Now Theorem 1.1 yields that $`^s\mathrm{\Psi }`$, as desired. ∎ ###### Proposition 4.2. All theorems of the paper hold for arbitrary $`FK`$. ###### Proof. Since the restrictions of the fundamental representations of a semisimple algebraic group over an algebraically closed field to relevant Chevalley groups over arbitrary subfields remain irreducible and can be realized over these subfields, only Theorems 1.1(ii) (or 3.3), 2.10, and 2.13 require some analysis. Let $`M`$ be the $`G_n`$-module $`\omega _i^{n+1}G_n`$ or $`W_i^n`$. First assume that $`F=\overline{F}`$ is algebraically closed. Set $`H=Sp_{2n}(F)`$. Let $`L`$ be the Lie algebra of $`H`$. For a root $`\alpha `$ of $`H`$ and $`tF`$ denote by $`x_\alpha (t)H`$ and $`X_\alpha L`$ the root elements in $`H`$ and $`L`$ associated with $`\alpha `$. It is well known that $`x_\alpha (t)(m)=(1+tX_\alpha )m`$ for $`mM`$ and long $`\alpha `$ (see, for instance, \[6, Lemma 1\]). For $`gG_n`$ set $`x_\alpha ^g(t)=gx_\alpha (t)g^1`$ and $`X_\alpha ^g=gX_\alpha g^1`$. It is clear that $`X_\alpha ^gL`$. It suffices to show that each $`G_n`$-submodule $`NM`$ is an $`H`$-submodule. Obviously, $`x_\alpha ^g(t)N=N`$ and $`X_\alpha ^gNN`$ for all long roots $`\alpha `$, $`gG_n`$, and $`tK`$. But this forces $`x_\alpha ^g(t)N=N`$ for all $`tF`$. However, using the commutator relations for the Chevalley groups of type $`C`$ (see, for instance, \[9, Lemma 15\]), one can deduce that the subgroup generated by all $`x_\alpha ^g(t)`$ with $`gG_n`$, $`tF`$, and long $`\alpha `$ coincides with $`H`$. (Here, in fact, it suffices to make computations within subgroups of type $`C_2`$ and show that our subgroup contains all short root subgroups). Hence $`N`$ is an $`H`$-module, as desired. Now let $`FK`$ be arbitrary. For a finite dimensional $`FG_n`$-module $`S`$ set $`\overline{S}=S_F\overline{F}`$ and denote the socle of $`S`$ by $`\mathrm{soc}(S)`$. Since $`dim\mathrm{Hom}_{FG_n}(E,S)=dim\mathrm{Hom}_{\overline{F}G_n}(\overline{E},\overline{S})`$ for any $`FG_n`$-module $`E`$, we have $`\mathrm{soc}(\overline{S})=\overline{\mathrm{soc}(S)}`$ if all composition factors of $`S`$ are absolutely irreducible. The same holds for other members of the socle series of $`S`$. If $`M=W_i^n`$, then $`M`$ and $`\overline{M}`$ are multiplicity-free and their submodules are completely determined by the sets of composition factors. Therefore the arguments on socles allow us to conclude that each submodule of $`\overline{M}`$ has the form $`\overline{S}`$ for some submodule $`SM`$. Let $`M=\omega _i^{n+1}G_n`$ with $`i,n1`$. Then the socle of $`M`$ contains a submodule $`V\omega _{i1}^n\omega _{i1}^n`$. Since $`M`$ and $`V`$ are selfdual and $`M`$ has only two composition factors isomorphic to $`\omega _{i1}^n`$, there exists a submodule $`D`$ of $`M`$ such that $`M=VD`$ and $`\overline{D}`$ is the unique submodule in $`\overline{M}`$ with $`\overline{M}=\overline{V}\overline{D}`$. Now one can see that the socle series of $`D`$ is determined by that of $`\overline{D}`$ and is described by Theorem 1.1(ii). ∎
warning/0003/cond-mat0003237.html
ar5iv
text
# Magnetoresistance in ordered and disordered double perovskite oxide, Sr2FeMoO6 ## I Introduction Colossal magnetoresistance (CMR) is a property that is of great technological potential, since the large change in the resistance (R) with the application of a magnetic field can be used effectively for mass storage magnetic devices. However, the well-known CMR materials, Mn based oxides, have significant CMR effect only at low temperatures and, therefore, are not suitable for room temperature applications. In their recent work, Kobayashi et al. have pointed out that the fully-ordered double perovskite Sr<sub>2</sub>FeMoO<sub>6</sub> with alternating Fe<sup>3+</sup> (3$`d^5`$, $`S=5/2`$) and Mo<sup>5+</sup> (4$`d^1`$, $`S=1/2`$) ferrimagnetically coupled ions exhibit substantial CMR even at room temperature. It is suggested that the half-metallic ferromagnetic (HMFM) state below T<sub>c</sub> is responsible for the magnetoresistance (MR) behavior via spin-dependent carrier scattering processes. While the spin-dependent scattering due to the intergrain tunneling effects enables one to construct manganite based devices , the intra-grain properties of the manganites are generally believed to be intimately connected with the intrinsic instability of Mn<sup>3+</sup> 3$`d^4`$ states arising from strong electron-phonon interaction as well as double-exchange interaction . Mo<sup>5+</sup> is also a Jahn-Teller ion with its 4$`d^1`$ $`{}_{}{}^{2}D`$ degenerate ground state. Thus, a priori it is not possible to rule out any contribution to the magnetoresistance of Sr<sub>2</sub>FeMoO<sub>6</sub> from effects other than intergrain spin-dependent carrier scatterings. Considering the complexity of the mechanism of magnetoresistance in the manganites, in which in addition to half-metallic ferromagnetism, various instabilities play important role, it is essential to address the factors responsible for the novel magnetoresistance properties of Sr<sub>2</sub>FeMoO<sub>6</sub>. Central to the HMFM state in Sr<sub>2</sub>FeMoO<sub>6</sub> as deduced by the band structure calculation is the long-range ordering of FeO<sub>6</sub> and MoO<sub>6</sub> octahedra alternatingly along three cubic axes. Disorder is expected to be detrimental to the half-metallic nature, giving rise to finite up- and down-spin densities of states at the Fermi energy, destroying the HMFM state; preliminary super-cell band structure calculations in our group indeed suggest that HMFM state is easily destroyed by a positional disorder at the Fe/Mo sites, though an overall ferrimagnetic order persists in the system. Keeping this in mind we have prepared for the first time Sr<sub>2</sub>FeMoO<sub>6</sub> by melt-quenching, where Fe and Mo sites are heavily disordered. A comparative study of disordered and ordered samples using x-ray diffraction, <sup>57</sup>Fe Mössbauer and the magnetoresistive properties allows us to delineate the effects arising from long range crystallographic order and those from factors other than half-metallic ferromagnetism in controlling the MR properties. ## II Experimental We prepared ordered Sr<sub>2</sub>FeMoO<sub>6</sub> following ref. 1 by hydrogen reduction of a mixture of SrCO<sub>3</sub>, MoO<sub>3</sub> and Fe<sub>2</sub>O<sub>3</sub>. The disordered sample was prepared by melt-quenching of a mixture of SrCO<sub>3</sub>, MoO<sub>3</sub>, Fe<sub>2</sub>O<sub>3</sub> and Mo-metal in an argon atmosphere. Mössbauer study was carried out in transmission geometry using a <sup>57</sup>Co/Rh matrix source. High resolution powder diffraction data were collected on a STOE STADI/P diffractometer. The resistivity measurements with and without an applied field were carried out using the standard four-probe dc measurements. ## III Results and discussion We show the powder x-ray diffraction patterns of the two samples in Fig. 1; Fig. 1(a) shows the pattern of the sample prepared by the conventional solid state techniques , while Fig. 1(b) shows the same for the melt-quenched sample prepared by the arc melting. The main diffraction peaks are the same in both the panels of Fig. 1 and these are readily ascribed to the perovskite lattice. The only difference between the two diffraction patterns is in terms of the existence of two weak peaks, at 19.6 and 38 in Fig. 1(a), but absence in Fig. 1(b). We have expanded the vicinity of these two angles in order to illustrate this point clearly in the respective panels. The existence of these two peaks in Fig. 1(a) establishes the presence of the supercell arising from the ordering of Fe and Mo sites alternatingly, giving rise to the double perovskite structure. Total or near-absence of these two order-related peaks in Fig. 1(b) suggests that the melt-quenched sample obtained from the arc furnace gives rise to heavy disrodering between the Fe and the Mo sites The extent of ordering at the Fe/Mo sites is easily estimated from x-ray diffraction (XRD) of ordered (A) \[Fig. 1(a)\] and disordered (B) \[Fig. 1(b)\] samples of Sr<sub>2</sub>FeMoO<sub>6</sub>. Rietveld refinement for both converged to the same cell dimensions ($`a`$ = $`b`$ = 5.566 Å and $`c`$ = 7.858 Å) with R$`{}_{wp}{}^{}`$ 8.0%. The ordering at the Fe and Mo sites in sample A was found to be $``$ 91%, slightly larger than the previously reported value of 87%. Near-absence of the long-range order-related peaks within the noise level of the data \[see insets I and II to Fig. 1(b)\] for sample B confirms the extensive disorder in this case. Rietveld analysis suggests approximately 31% ordering in this case. The effect of disorder at the Fe-sites is also probed by <sup>57</sup>Fe Mössbauer spectra of these two samples at 300 K and 4.2 K (Fig. 2); a comparison of the raw spectra for the two samples at these temperatures is sufficient to establish the relative degrees of disorder in site occupancies. The spectra for the sample B exhibits relatively larger line-widths in contrast to those of sample A, implying a larger disorder in sample B. All the spectra are magnetic hyperfine split, indicating magnetic ordering of the materials. The spectrum at 300 K for sample A favorably compares with that reported earlier . Our analysis of the data at 4.2 K for sample A yields a saturation hyperfine field of about 480 kOe, with a small distribution of about 10 kOe. This value is somewhat smaller than that known for Fe<sub>2</sub>O<sub>3</sub>. The hyperfine field for sample B has a much larger distribution (490 $`\pm `$ 60 kOe), quantitatively confirming a strong variation in the chemical environment for Fe from site to site. The chemical composition was however found to be identical between the two samples by energy dispersive x-ray analysis. Moreover, the grain size and morphology were also found to be similar between the two samples by scanning electron microscopy (SEM). The above results establish that it is indeed possible to obtain highly ordered and extensively disordered samples with respect to Fe and Mo occupancies without otherwise disturbing the structural integrity of this double perovskite system. We show the magnetoresistance, $`MR(H,T)=\{R(H,T)R(0,T)\}/R(0,T)`$, of ordered and disordered samples as a function of applied magnetic field, H, at 300 K and 4.2 K in Fig 3; in the same figure we reproduce the results from Kobayashi et al. . At both the temperatures, the ordered sample is characterised by a sharp magnetoresistive response in the low-field region, though the magnitude of the magnetoresistance is considerably larger at the lower temperature, in agreement with the published literature . The present ordered sample has a sharper low-field response and a larger magnitude, particularly at the low temperature compared to the previous result . This improved magnetoresistance response is possibly related to somewhat higher degree of ordering in the present sample. This improvement in the low magnetic field response suggests that a fully (100%) ordered sample is likely to be a very good candidate for real-life device applications. Beyond about 1 T, the magnetoresistance of the ordered sample exhibits a slower change at higher field strengths, without showing any signs of saturation upto the highest magnetic field probed in the present experiments. The MR changes significantly (by about 6.5% at 4.2 K and 3% at 300 K) in the larger field regime between 1 and 7 T. In contrast, the MR of the disordered sample does not exhibit the low-field sharp magnetic response below 1 T. Since the two samples differ only in the extent of Fe/Mo ordering, the low-field rapid variation in MR of sample A evidently arises from the long-range order, leading to the half metallic ferromagnetic state of sample A and the consequent strong intergrain spin-dependent scattering. It is, however, intriguing to note that there is a substantial negative magnetoresistance in the disordered sample and the high field behaviors of the MR in the two samples are similar, the two curves being approximately parallel in the high-field region at both the temperatures shown in Fig. 3. This implies that there is a common intragrain origin of this high-field MR behavior in the disordered and the ordered samples. We have carried out band structure calculations, simulating the disordered system within a supercell approach where the Fe and Mo atoms were allowed to occupy the other sites in contrast to the perfectly ordered system. Preliminary results from these calculations show that in the disordered state, both up- and down-spin states exist at the Fermi energy; this indicates that the half-metallic ferromagnetic ground state is not realised for any disordering of the Fe and the Mo sublattices, though the system remains to be a ferrimagnet. Thus, the high-field, intragrain contribution to the negative magnetoresistance is independent of the long range ordering of Fe and Mo atoms and the cosnequent existence of the HMFM state. Instead, it is to be associated with the usual negative magnetoresistance observed for ferromagnetic substances, arising from a suppression of the spin-fluctuations with the application of an external magnetic field. In conclusion, we have prepared highly ordered and disordered samples of the perovskite oxide, Sr<sub>2</sub>FeMoO<sub>6</sub>. We characterised the extent of ordering by x-ray diffraction and Mössbauer studies of these samples. Rietveld analysis of the powder diffraction pattern shows that the extent of ordering in the Fe/Mo sites is about 91% for the ordered sample, while it is about 31% for the disordered sample. We point out that the disordered sample is not a half-metallic ferromagnet, though it is magnetic even at the room temperature. We showed that the magnetoresistance of the two samples are distinguished by the existence of the sharp changes in the magnetoresistance at low fields for the ordered sample. In the case of the disordered sample, there is no such sharp changes in the MR, though a negative MR is seen at all fields for the disordered sample also. The MR of the disordered sample is very similar to that of the ordered sample in the high-field regime, indicating a common origin. Thus, we conclude that the low-field magnetoresistance is dominated by the intergrain spin-dependent scattering of the highly spin polarized charge-carriers of the half-metallic ferromagnet, the high-field negative magnetoresistance behaviours in both the samples are intrinsic, intragrain property and arise from the suppression of spin-fluctuations under the application of a magnetic field on a magnetic system. Acknowledgements: The authors thank Prof. B. Sriram Shastry and Dr. Tanusri Saha Dasgupta for useful discussions. ## IV figure captions Fig. 1. X-ray diffraction patterns of (a) ordered and (b) disordered Sr<sub>2</sub>FeMoO<sub>6</sub>. Order-related peaks, appear at 19.6 and 38, shown in the insets on an expanded scale. Fig. 2. <sup>57</sup>Fe Mössbauer spectra of the disordered and ordered Sr<sub>2</sub>FeMoO<sub>6</sub> at 300 and 4.2 K. Fig. 3. The comparisons of percentage magnetoresistance between the ordered and disordered samples as well as the result presented by Kobayashi et al. at (a) 4.2 K and (b) 300 K.
warning/0003/hep-th0003277.html
ar5iv
text
# Abstract ## Abstract A new model of supersymmetry between bosons and fermions is proposed. Its representation space is spanned by states with $`𝒫𝒯`$ symmetry and real energies but the inter-related partner Hamiltonians themselves remain complex and non-Hermitian. The formalism admits vanishing Witten index. ## 1 Introduction Supersymmetry (SUSY, ) offers a chance of unification of bosons with fermions in various branches of physics . Mathematically, it mixes the commutators and anticommutators in a single (so called graded) Lie algebra. In its simplest form sl(1/1) this symmetry algebra is generated by two supercharges $`𝒬`$, $`\stackrel{~}{𝒬}`$ and a Hermitian Hamiltonian $``$. These three generators are related by the anticommutation rules $$\{𝒬,\stackrel{~}{𝒬}\}=,\{𝒬,𝒬\}=\{\stackrel{~}{𝒬},\stackrel{~}{𝒬}\}=0$$ (1) and commutativity $$[,𝒬]=[,\stackrel{~}{𝒬}]=0.$$ (2) SUSY finds an enormous appeal in particle physics and field theory . In this application, unfortunately, there exists the strong conflict between the theory and experiment. Due to the continuing absence of observation of any bosonic-fermionic degenerate multiplets, any form of SUSY must be badly broken as a consequence . Attempts to resolve the latter physical problem encounter nontrivial mathematical difficulties . Possible mechanisms of SUSY breaking are currently being exposed to an intensive research . One of the most feasible ways of their analysis is offered by the representation of SUSY algebras in a zero-dimensional field theory, i.e., in quantum mechanics . We intend to contribute to this effort by weakening certain assumptions concerning, first of all, the hermiticity of $`H`$. ## 2 PT symmetric non-Hermitian models Conventional supersymmetric quantum mechanics does not immediately admit complex potentials. Only recently, the first attempts in this direction have been made . A significant improvement of our understanding of the underlying complex dynamics has been offered by Bender and Boettcher . Within their so called $`𝒫𝒯`$ symmetric quantum mechanics they proposed a replacement of the usual hermiticity of the Hamiltonian $`H`$ by its mere commutativity with the product of the parity $`𝒫`$ and the time reversal $`𝒯`$, $$H𝒫𝒯=𝒫𝒯H.$$ It has been shown by several methods that one may get real energy spectrum in many different systems of this type . Similar analytic assumptions made in connection with quantized systems is not so unusual in the mathematically oriented literature . Its appeal in physical applications is also undeniable and ranges from perturbative and semiclassical methods and considerations up to the practical computation of resonances . In such a setting, the construction of representations of SUSY algebras encounters several new challenges. Some of them will be addressed in what follows. ### 2.1 Facilitated normalization For a concise exposition of some of the related open questions let us first recall a quartic partially solvable potential $$V_1(r)=r^4+2ir$$ of ref. . Its one-dimensional Schrödinger equation $$\frac{d^2}{dr^2}\mathrm{\Psi }(r)+V_1(r)\mathrm{\Psi }_1(r)=E_1\mathrm{\Psi }_1(r),r(\mathrm{},\mathrm{})$$ possesses a formal solution at zero energy $`E_1=0`$. Forgetting, for the time being, that this solution is not normalizable in the usual sense, $$\mathrm{\Psi }_1^{(0)}(r)=\mathrm{exp}\left(\frac{ir^3}{3}\right)L_2(\mathrm{},\mathrm{})$$ (3) we can construct its superpotential $$W_1(r)=\left[\frac{d}{dr}\mathrm{\Psi }_1^{(0)}(r)\right]/\mathrm{\Psi }_1^{(0)}(r)=ir^2$$ and derive formally the supersymmetric partner potential $$V_2(r)=W_1^2(r)+W_1^{}(r)=W_2^2(r)W_2^{}(r)=r^42ir,r(\mathrm{},\mathrm{})$$ as well as the parallel ground-state-like solution $$\mathrm{\Psi }_2^{(0)}(r)=\mathrm{exp}\left(\frac{ir^3}{3}\right)L_2(\mathrm{},\mathrm{}).$$ (4) Obviously, in such a model the formal SUSY transformation $`12`$ degenerates to the mere time reversal represented, for our present purposes, by the above-mentioned operator $`𝒯`$ which replaces $`i`$ by $`i`$ , $$𝒯\mathrm{\Psi }_2^{(0)}(r)=\mathrm{\Psi }_1^{(0)}(r),𝒯V_2(r)𝒯=V_1(r).$$ Within the less naive framework of the $`𝒫𝒯`$ symmetric quantum mechanics the latter example proves better understood. Firstly, in the light of the analyticity of our model we can restore the normalizability of its wave functions (3) and (4) by suitable deformations of the coordinate axis in complex plane . This can be achieved by the mere shifts $$r=r_{1,2}(x)=x\pm i\epsilon ,\epsilon >0,x(\mathrm{},\mathrm{})$$ of the respective integration paths. This guarantees that the wave functions become asymptotically vanishing as required, $`\mathrm{\Psi }_j^{(0)}[r_j(\pm \mathrm{})]0`$. Unfortunately, we have to pay a high price. After one verifies that $$V_2=W_2^2W_2^{}=x^4+4i\epsilon x^3+\mathrm{}W_1^2+W_1^{}=x^44i\epsilon x^3+\mathrm{},$$ we have to conclude that our two new, $`𝒫𝒯`$ symmetrized interactions $`V_{1,2}`$ cease to be inter-related by a supersymmetry. In what follows we intend to re-solve the puzzle. In essence, we shall generalize the original Witten’s quantum mechanical construction . Our attention will be paid to situations where the above-exemplified loss of a SUSY partnership could be re-established anew. In brief, we shall propose an entirely new representation of the supersymmetric algebra within the framework of the $`𝒫𝒯`$ symmetric quantum mechanics. ### 2.2 A toy model without SUSY We intend to introduce our proposal via a few explicit examples. Particular attention will be paid to the two manifestly $`𝒫𝒯`$ symmetric potentials $$V^{()}(x)=4i(xi\epsilon )(xi\epsilon )^4,$$ (5) $$V^{(+)}(x)=\frac{2}{(x+i\epsilon )^2}(x+i\epsilon )^4.$$ (6) Their doublet resembles the previous pair by the similar choice of the respective domains $`r_{(\pm )}(x)=x\pm i\epsilon `$. Conveniently, both their shifts are equal and given by the same positive constant $`\epsilon >0`$. Our new examples (5) and (6) also exhibit the so called quasi-exact solvability revealed in refs. and , respectively. Meaning just that a few exact bound states remain at our disposal in an elementary form , this property offers us the two exact zero-energy bound-state solutions $$\psi ^{()}(x)=(xi\epsilon )\mathrm{exp}\left(i\frac{(xi\epsilon )^3}{3}\right)L_2(\mathrm{},\mathrm{}),$$ $$\psi ^{(+)}(x)=\frac{1}{x+i\epsilon }\mathrm{exp}\left(+i\frac{(x+i\epsilon )^3}{3}\right)L_2(\mathrm{},\mathrm{})$$ representing, presumably, ground states. Both these wave functions are bounded and normalizable if and only if their common real parameter $`\epsilon `$ is positive. This is similar to our previous illustration while, in contrast, the new superpotentials $$W^{(\pm )}(x)=\left[\frac{d}{dx}\psi ^{(\pm )}(x)\right]/\psi ^{(\pm )}(x)=\pm \left[\frac{1}{x\pm i\epsilon }i(x\pm i\epsilon )^2\right].$$ (7) differ more than just by an overall sign. In the unphysical extreme of the vanishing parameter $`\epsilon 0`$ we would arrive at the standard SUSY connecting our two Hamiltonians $`H^{(\pm )}`$ but no serious progress seems to have been achieved. We again “stumble” over the normalizability of our wave functions which would be lost in the SUSY limit. Still, there is a difference. We are going to show below that our new doublet of models can be supersymmetrized after one modifies the usual recipe. ## 3 PT symmetric supersymmetry A key observation of our present proposal is that many $`𝒫𝒯`$ symmetric systems are defined off the real axis. Boundary conditions $`lim_{|r|\mathrm{}}\psi (r)=0`$ are in general located within wedges bounded by Stokes’ lines . Locally, the paths of integration can be deformed whenever necessary. In contrast to our introductory examples $`V_{1,2}`$, the mere “time-reflection” conjugation $`𝒯`$ itself does not now map our new Hamiltonians $`H^{(\pm )}`$ upon each other. Still, an active use of $`𝒯`$ will be a key ingredient in our forthcoming construction. ### 3.1 Innovated factorization of Hamiltonians With our functions (7) taken just as a particular illustration, let us now assume their arbitrary form and introduce the four related operators $$A^{(\pm )}=\frac{d}{dx}+W^{(\pm )}(x),B^{(\pm )}=\frac{d}{dx}+W^{(\pm )}(x)$$ which induce the traditional Riccati-equation formulae $$H^{(\pm )}=B^{(\pm )}A^{(\pm )}=_x^2+[W^{(\pm )}(x)]^2\left[W^{(\pm )}(x)\right]^{}.$$ We have to fit these two Hamiltonians into a generalized SUSY scheme of the type (1) + (2). In the first step, we explored reordered products. Returning to our explicit examples $`V^{(\pm )}`$ for inspiration, we did not succeed in the $`{}_{}{}^{(+)}`$superscripted case at all. Fortunately, in the second case we were able to verify by immediate insertions the strict validity of the only slightly nonstandard rule $$H^{(+)}=𝒯A^{()}B^{()}𝒯.$$ (8) This formula is, in a way, our central point. Indeed, once we arrange our doublet of Hamiltonians into the following two-dimensional array $$=\left[\begin{array}{cc}H^{()}& 0\\ 0& H^{(+)}\end{array}\right]=\left[\begin{array}{cc}B^{()}A^{()}& 0\\ 0& 𝒯A^{()}B^{()}𝒯\end{array}\right]$$ we recover immediately all the necessary SUSY rules (1) and (2), provided only that we introduce the following modified representation of the supercharges, $$𝒬=\left[\begin{array}{cc}0& 0\\ 𝒯A^{()}& 0\end{array}\right],\stackrel{~}{𝒬}=\left[\begin{array}{cc}0& B^{()}𝒯\\ 0& 0\end{array}\right].$$ (9) In contrast to the usual SUSY constructions our new supercharges are not correlated by any Hermitian conjugation anymore. This is our main methodological gain. A new hope is created that some “no-go” theorems of the traditional Hermitian theories could be overcome within our new SUSY framework. A key technical difficulty with this hope lies in its dependence on the specific choice of our example. Fortunately, one can re-analyze our fundamental re-arrangement in the more general context where the explicit form $$\left[W^{(+)}\right]^2\left[W^{(+)}\right]^{}=\left\{\left[W^{()}\right]^2+\left[W^{()}\right]^{}\right\}^{}$$ of eq. (8) may be called a ladder equation. It glues superpotentials in the language of the higher order SUSY . An explicit solution of the latter equation exists and can be expressed in the parametric form using an arbitrary complex function $`f(x)`$, $$W^{(+)}=\frac{f^{}}{2f}f,W^{()}=\frac{f^{}}{2f^{}}f^{}$$ (10) It is worth noticing that with the shape invariant $`f(x)=i\lambda \mathrm{sech}\mu x`$, one re-discovers the amazing though very special relationship between the non-Hermitian and Hermitian exactly solvable models of ref. . In the latter class of examples with the purely imaginary $`f(x)`$ our “new SUSY” coincides with the ordinary, “classical SUSY” since the lower Hamiltonian remains real, $`H^{(+)}=𝒯H^{(+)}𝒯`$. With arbitrary $`f(x)`$ equation (10) suggests that a certain modified Darboux transformation is at work for $`H^{(+)}𝒯H^{(+)}𝒯`$. In this way, also another, “fully complex” illustration $`f(x)=i(x+i\epsilon )^2`$ returns us back to our “new-SUSY” example $`V^{(\pm )}`$ which manifestly breaks the standard SUSY at $`\epsilon >0`$. ### 3.2 Intertwining relations Much in the same spirit as when one seeks possible interwinings of operators in the context of higher-order SUSY let us now return to their present realization, starting from the assumption of reality of the energies in the Schrödinger equation $$H^{()}\psi _n^{()}(x)=B^{()}A^{()}\psi _n^{()}(x)=E_n^{()}\psi _n^{()}(x).$$ In the light of its possible re-factorization (8) let us now pre-multiply it by a suitable operator from the left, $$[𝒯A^{()}][B^{()}𝒯][𝒯A^{()}\psi _n^{()}(x)]H^{(+)}\psi _m^{(+)}(x)=E_n^{()}[𝒯A^{()}\psi _n^{()}(x)].$$ This correspondence can be accompanied by the second Schrödinger equation $$H^{(+)}\psi _m^{(+)}(x)=B^{(+)}A^{(+)}\psi _m^{(+)}(x)=E_m^{(+)}\psi _m^{(+)}(x).$$ A comparison results in the general relationship $$\psi _m^{(+)}(x)=𝒯A^{()}\psi _n^{()}(x),E_m^{(+)}=E_n^{()}.$$ In parallel, we can also re-write $`H^{()}\psi _k^{()}(x)`$ in the re-factorized form $$[B^{()}𝒯]𝒯A^{()}[B^{()}𝒯\psi _m^{(+)}(x)]=E_m^{(+)}[B^{()}𝒯\psi _m^{(+)}(x)]E_k^{()}\psi _k^{()}(x)$$ and deduce that $$\psi _k^{()}(x)=B^{()}𝒯\psi _m^{(+)}(x),E_k^{()}=E_m^{(+)}.$$ One has to be careful with a quick assignment of the labels. In general, one cannot be sure about the ordering of the levels. In accord with several explicit examples their various permutations could occur here in general. Fortunately, many rules concerning the ordering of levels in real potentials find their direct analogues in the complexified Sturm Liouville oscillation theorems . Moreover, their “almost standard” form applies in the case of the present “asympotically almost real” examples (5) and (6). For them it is possible to show that $`m=n=k`$. In such a case the current rules of reconstruction of the partner spectra (the variety of constructive examples of which can be found in refs. ) can be restored in their full strength. ## 4 Discussion We have seen that a core of applicability of our new form of the SUSY transformation to a complex force lies in our understanding of the normalizability of the wave functions in the $`𝒫𝒯`$ symmetric formalism. Still, even in this context many formal questions remain open. For example, due to a spontaneous breakdown of the $`𝒫𝒯`$ symmetry the energies can sometimes coalesce in the complex conjugated pairs . The rigorous foundations of the reality of the spectra must be always scrutinized anew. Mathematically, a key feature of the present construction lies in a difference in the arguments $`x\pm i\epsilon =r_{(\pm )}(x)`$ of our model potentials, i.e., in the domains of definition of the Hamiltonians. This freedom admits a further generalization to all the integrations paths $`r_{(\pm )}(x)`$ which are coupled by the reflection with respect to the real axis or, in the present language, by the time reversal. In this sense the operator $`𝒯`$ plays a double role: To its original meaning of a reflection of the complex plane (for coordinates) one has to add its use in our innovated hypercharges and in the alternative factorization of $`H^{(+)}`$. The relevance of the present proposal is enhanced by several unexpected observations. Firstly we notice that our supersymmetric mapping does not seem to require the current complementary comment about the missing partner of the zero energy bound state itself. This means that the Witten index vanishes, $`n_Bn_F=0`$. Still, due to the broken hermiticity of our Hamiltonians, both zero-energy states remain normalizable so that the supersymmetry itself remains unbroken. We may conclude that the present $`𝒫𝒯`$ symmetric formalism is quite different from its current Hermitian predecessors. It resembles the models with periodic potentials and the higher order SUSY quantum mechanics in the irreducible case. Let us recall that in the latter context one also cannot express the intermediate Hamiltonian as a Hermitian operator . Undoubtedly, the immediate relationship to the vanishing Witten index makes our construction very appealing. In a summary, we could now distinguish between the three different forms of SUSY. Firstly, one defines the standard one in a formalism using the real potentials and superpotentials . Secondly, a use of the complex superpotentials and charges which are not Hermitian conjugate of each other forms simply an opposite extreme . Thirdly, our present formalism stays somewhere in between. It constrains the latter unrestricted freedom by the fairly nontrivial $`𝒫𝒯`$ symmetry but, in contrast to the current Witten’s SUSY it is not restricted to non-negative Hamiltonian operators. ## Acknowledgements M. Z. acknowledges the hospitality of INFN and the University of Bologna and also the grant Nr. A 1048004 of GA AS CR. B. B. thanks Prof. C. Quesne for several fruitful discussions.
warning/0003/nucl-th0003037.html
ar5iv
text
# Theory of doorway states for one-nucleon transfer reactions. II. Model-independent study of nuclear correlation effects ## 1 Introduction The nucleon-nucleon interaction still remains one of the central problems of the nuclear structure theory. It seems natural to use the free-space forces, but the most of the nuclear structure theorists prefer the effective ones instead. The basic motivations are as follows. (a) Historically the first one arose from the wide-spread belief in sixties that the free-space $`NN`$ potential has a hard repulsive core. Clearly such interaction does not apply directly because of the divergence of finite-order Feynman diagrams. The most popular way to overcome this difficulty is to calculate the Brueckner $`G`$-matrix and use it as the effective interaction in the Hartree-Fock problem. This is the Hartree–Fock–Brueckner approximation, see and the references therein for details. But soon it became clear that the description of the two-nucleon system, i.e. the deuteron properties and the elastic $`NN`$ scattering phase shifts below the pion production threshold, does not require the hard repulsive core. As a result all the contemporary $`NN`$ potentials are of soft core character, and so the above difficulty does not really exist. In such conditions the calculation of the Brueckner $`G`$-matrix is not compulsory. (b) The renormalization of the free-space interaction due to the medium polarization effects. But such effects are treated by conventional methods of quantum many-body theory, and therefore the above reason is not the argument for the effective forces. (c) The medium QCD-renormalization due to the fact that the quark composition of the QCD vacuum is changed in nuclear medium thus leading to the changes of both the mesons and the meson–nucleon vertices . Such processes can hardly be described in all details since the exact theory does not yet exist for the nonperturbative region. Nevertheless, at least one exact statement can be done. The QCD-renormalized interaction is the functional of nuclear density possessing the following obvious property: it turns into the free-space one in the zero-density limit. For this reason it can be represented as the following functional expansion: $`f_{QCDr}(𝐫_1𝐫_2\{\rho \})`$ $`=`$ $`f_2(|𝐫_1𝐫_2|)+{\displaystyle f_3(|𝐫_1𝐫_2|,|𝐫_1𝐫^{}|)\rho (r^{})d^3𝐫^{}}+`$ (1) $`+`$ $`{\displaystyle f_4(|𝐫_1𝐫_2|,|𝐫_1𝐫^{}|,|𝐫_1𝐫^{\prime \prime }|)\rho (r^{})\rho (r^{\prime \prime })d^3𝐫^{}d^3𝐫^{\prime \prime }}+\mathrm{}`$ In this way we conclude that the medium QCD-renormalization is equivalent to the existence of many-particle $`NN`$ forces in addition to the two-particle ones. This conclusion is confirmed by the fact that the physics of strong interaction is essentially nonlinear. But the nonlinearity automatically leads to many-particle forces. The above reasons are grounds for our starting point: both the two-particle and many-particle $`NN`$ forces must be taken into account for the treatment of nuclear structure. The three-particle $`NN`$ forces are indeed included for the calculations of few-nucleon systems . But the following questions arises: is this sufficient for complex nuclei ? This point as well as a number of those concerning the nuclear structure might be elucidated if the model-independent object would exist in nuclei. Our approach is based on the fact that such objects do really exist. As discussed in our previous work they are the doorway states for one-nucleon transfer reaction. We showed that (i) at least the three-particle repulsion and the four-particle attraction must be taken into account in addition to the two-particle forces; (ii) the nuclear relativity is really existing phenomenon rather than the hypothesis of J.D. Walecka ; (iii) the dominant contribution to the isovector nuclear potential is provided by the many-particle forces; (iv) the observed spectra of the doorway states can be used to specify the neutron density distributions in nuclei; such specified densities are indeed obtained for the closed-shell nuclei <sup>40</sup>Ca, <sup>90</sup>Zr and <sup>208</sup>Pb. In the present work our approach is applied to the ”empirical” studies of nuclear correlation effects owing to which the nuclear wave functions are different from the Slater determinants. Such effects are treated by a variety of approximate methods since the exact ones do not exist. For this reason it is very important to get a model-independent quantitative information about the above effects. The possibility provided by our approach is based on the fact that the single-particle states of nucleon in nuclear static field (which are just the doorway states for one-nucleon transfer reactions ) are correlation-free objects in contrast to the Landau–Migdal quasiparticles (which include the correlations by definition) and single-particle states in nuclear Hartree–Fock calculations with effective forces (where the correlations are implicitly included in the phenomenological effective force parameters). So, calculating the correlation-free quantities and comparing them with the observed ones we get a quantitative measure of the correlation effects. In Sect.2 this procedure is applied to the nucleon density distributions whereas the problems concerning the nuclear binding energy are discussed in Sect.3. ## 2 Density distributions The nucleon density distribution in the ground state of nucleus $`A`$ is (see for the notations) $`\rho (r)`$ $`=`$ $`A_0|\psi ^+(x)\psi (x)|A_0={\displaystyle \underset{j}{\overset{(A1)}{}}}A_0|\psi ^+(x)|(A1)_j(A1)_j|\psi (x)|A_0=`$ (2) $`=`$ $`{\displaystyle \underset{j}{\overset{(A1)}{}}}\mathrm{\Psi }_j^+(x)\mathrm{\Psi }_j(x)={\displaystyle _C}{\displaystyle \frac{d\epsilon }{2\pi i}}G(x,x;\epsilon ).`$ The integration contour $`C`$ includes the real axis and the infinite radius semicircle in the upper part of the complex $`\epsilon `$ plane. As seen from Eq.(1) the density is expressed through the single-particle amplitudes of the $`(A1)`$ nuclear states. Expanding them over the complete set of the doorway states, $$\mathrm{\Psi }_j(x)=\underset{\lambda }{}C_j^{(\lambda )}\psi _\lambda (x)$$ (3) and putting Eq.(3) into Eq.(2) we get $$\rho (r)=\underset{\lambda }{}n_\lambda |\psi _\lambda (x)|^2+2\underset{\genfrac{}{}{0pt}{}{\lambda \nu }{\nu >\lambda }}{}\rho _{\lambda \nu }\psi _\lambda ^+(x)\psi _\nu (x),$$ (4) where $$n_\lambda =\rho _{\lambda \lambda }=\underset{j}{\overset{(A1)}{}}|C_j^{(\lambda )}|^2=\underset{j}{\overset{(A1)}{}}s_j^{(\lambda )},\rho _{\lambda \nu }=\underset{j}{\overset{(A1)}{}}C_j^{(\lambda )^{}}C_j^{(\nu )}$$ (5) (actually the coefficients $`C_j^{(\lambda )}`$ are real quantities if the parity violation due to weak interaction is disregarded). The diagonal elements $`\rho _{\lambda \lambda }=n_\lambda `$ are the doorway state occupation numbers. Indeed, the particle number is $$N=\rho (r)d^3𝐫=\underset{\lambda }{}n_\lambda ,$$ (6) since the nondiagonal elements do not contribute because of the orthogonality between $`\psi _\lambda `$ and $`\psi _\nu `$. The quantities $`n_\lambda `$ and $`\rho _{\lambda \nu }`$ obey the following limitations: $$0<n_\lambda <1,|\rho _{\lambda \nu }|<\frac{1}{2}(n_\lambda +n_\nu ).$$ (7) The first follows from the fact that the doorway states are distributed over the actual ones of both the $`(A1)`$ and $`(A+1)`$ nuclei, see Eq.(13c) in Ref., whereas the second is a consequence of the Cauchy–Bunjakowsky inequality. The facts $`n_\lambda <1`$ and $`\rho _{\lambda \nu }0`$ reflect the Fermi surface smearing due to the correlation effects. The quantities $`n_\lambda `$ and $`\rho _{\lambda \nu }`$ carrying the quantitative information about these effects can be found ”empirically” by considering the relation (4) together with the limitations (7) as equations for $`n_\lambda `$ and $`\rho _{\lambda \nu }`$. The results will be published elsewhere. As a result of the correlations a part of nucleons is out of the nuclear Fermi-surface. The number $`N_{out}`$ of such nucleons is calculated by comparing the observed density distribution with the correlation-free one $$\rho _{cf}(r)=\underset{\lambda }{}\mathrm{\Theta }(\epsilon _F\epsilon _\lambda )|\psi _\lambda (x)|^2.$$ (8) $`\epsilon _F`$ is the Fermi level energy. Such comparison is shown in Fig.1 for the proton density distribution in <sup>40</sup>Ca. As seen from the figure the correlation-free density contains more nucleons in the inner region $`0<r<r_i`$ than the observed one, $`r_i`$ is the intersection point. The situation in the outer region $`r_i<r<\mathrm{}`$ is clearly opposite since both densities correspond to the same number of nucleons. The number of redistributed nucleons is $$N_{out}=4\pi \underset{0}{\overset{r_i}{}}[\rho _{cf}(r)\rho (r)]r^2𝑑r=4\pi \underset{r_i}{\overset{\mathrm{}}{}}[\rho (r)\rho _{cf}(\rho )]r^2𝑑r.$$ (9) $`r_i`$ is the intersection point: $`\rho _{cf}(r_i)=\rho (r_i)`$. This is just the $`N_{out}`$ because the only reason for the redistribution is the Fermi-surface smearing due to the correlations. The $`N_{out}`$ numbers in doubly closed-shell nuclei are shown in Table 1. As seen from the table these nucleons are a rather appreciable part of the total mass number. To our knowledge this fact was first mentioned by Frankfurt and Strikman on the basis of the analysis of high-momentum components (i.e. those with $`k>300`$MeV/c) of the nucleon momentum distributions in nuclei. According to their most recent results for these data the $`A_{out}/A`$ ratio is $`(20\pm 3)\%`$ for heavy nuclei. This is in reasonable agreement with our results. It is worth mentioning that our calculations do not tell anything about the nature of the underlying correlations whereas the high-momentum tales of the momentum distributions arise from the $`NN`$ interactions at short distances . Therefore the reasonable agreement between the two results gives rise to the conclusion that the main reason for the Fermi-surface smearing in doubly closed-shell nuclei is due to the short-range correlations. ## 3 Binding energy We also can calculate the part of the nuclear binding energy which is caused by the motion of nucleons in nuclear static field. Such static energy partly includes the correlation effects since it is expressed through the observed nucleon density distributions. The comparison between the observed binding energy $`_b`$ and the static one $`_{st}`$ gives the measure of the proper correlation energy of nucleus. To make this point clear let us derive the exact expression for the binding energy. Following the procedure of Ref. we get $`_b={\displaystyle }dx{\displaystyle \underset{c}{}}{\displaystyle \frac{d\epsilon }{2\pi i}}tr\widehat{k}_xG(x,x;\epsilon )+{\displaystyle \frac{1}{2}}{\displaystyle }{\displaystyle }dxdx_1{\displaystyle }{\displaystyle \underset{c}{}}{\displaystyle \frac{d\epsilon d\epsilon _1}{(2\pi i)^2}}f_2(|𝐫𝐫_1|);\omega )K_2(x,x_1;x,x_1;\epsilon ,\epsilon _1)+`$ $`+`$ $`{\displaystyle \frac{1}{3}}{\displaystyle 𝑑x𝑑x_1𝑑x_2\underset{c}{}\frac{d\epsilon d\epsilon _1d\epsilon _2}{(2\pi i)^3}f_3(|𝐫𝐫_1|,|𝐫𝐫_2|;\omega ,\omega _1)K_3(x,x_1,x_2;x,x_1,x_2;\epsilon ,\epsilon _1,\epsilon _2)}+`$ $`+`$ $`{\displaystyle \frac{1}{4}}{\displaystyle }{\displaystyle }{\displaystyle }{\displaystyle }dxdx_1dx_2dx_3{\displaystyle }{\displaystyle }{\displaystyle }{\displaystyle \underset{c}{}}{\displaystyle \frac{d\epsilon d\epsilon _1d\epsilon _2d\epsilon _3}{(2\pi i)^4}}f_4(|𝐫𝐫_1|,|𝐫𝐫_2|,|𝐫𝐫_3|;\omega ,\omega _1,\omega _2)\times `$ $`\times `$ $`K_4(x,x_1,x_2,x_3;x,x_1,x_2,x_3;\epsilon ,\epsilon _1,\epsilon _2,\epsilon _3),`$ (10) where $`K(x_1,\mathrm{}x_n;x_1^{}\mathrm{}x_n^{};\epsilon _1\mathrm{}\epsilon _n)`$ are the $`n`$-particle Green functions. We accounted for the three- and four-particle forces in addition to the two-particle ones as well as the possible dependence of the interactions on the appropriate energy transfers $`\omega _i`$. In these terms the single-particle Green function is $`(\epsilon \widehat{k}_x)G(x,x^{};\epsilon )=\delta (xx^{})+{\displaystyle 𝑑x_1\underset{c}{}\frac{d\epsilon _1}{2\pi i}f_2(|𝐫𝐫_1|;\omega )K_2(x,x_1;x^{},x_1^{};\epsilon ,\epsilon _1)}+`$ $`+`$ $`{\displaystyle 𝑑x_1𝑑x_2\underset{c}{}f_3(|𝐫𝐫_1|,|𝐫𝐫_2|;\omega ,\omega _1)K_3(x,x_1,x_2;x^{},x_1,x_2;\epsilon ,\epsilon _2)}+`$ $`+`$ $`{\displaystyle }{\displaystyle }{\displaystyle }dx_1dx_2dx_3{\displaystyle }{\displaystyle \underset{c}{}}{\displaystyle }{\displaystyle \frac{d\epsilon _1d\epsilon _2d\epsilon _3}{(2\pi i)^3}}f_4(|𝐫𝐫_1|,|𝐫𝐫_2|,|𝐫𝐫_3|;\omega ,\omega _1,\omega _2)\times `$ $`\times `$ $`K_4(x,x_1,x_2,x_3;x^{},x_1,x_2,x_3;\epsilon ,\epsilon _1,\epsilon _2,\epsilon _3).`$ (11) On the other hand the Dyson equation is $$(\epsilon \widehat{k}_x)G(x,x^{};\epsilon )=\delta (xx^{})+𝑑x_1M(x,x_1;\epsilon )G(x_1,x^{};\epsilon ),$$ (12) and therefore $`{\displaystyle }M(x,x_1;\epsilon )G(x_1,x^{};\epsilon )dx_1={\displaystyle }dx_1{\displaystyle \underset{c}{}}{\displaystyle \frac{d\epsilon _1}{2\pi i}}f_2(|𝐫𝐫_1);\omega )K_2(x,x_1;x^{},x_1;\epsilon ,\epsilon _1)+`$ $`+`$ $`{\displaystyle 𝑑x_1𝑑x_2\underset{c}{}\frac{d\epsilon _1d\epsilon _2}{(2\pi i)^2}f_3(|𝐫𝐫_1|,|𝐫𝐫_2|;\omega ,\omega _1)K_3(x,x_1,x_2;x^{},x_1,x_2;\epsilon ,\epsilon _1,\epsilon _2)}+`$ $`+`$ $`{\displaystyle }{\displaystyle }{\displaystyle }dx_1dx_2dx_3{\displaystyle }{\displaystyle \underset{c}{}}{\displaystyle }{\displaystyle \frac{d\epsilon _1d\epsilon _2d\epsilon _3}{(2\pi i)^3}}f_4(|𝐫𝐫_1|,|𝐫𝐫_2|,|𝐫𝐫_3|;\omega ,\omega _1,\omega _2)\times `$ $`\times `$ $`K_4(x,x_1,x_2,x_3;x^{},x_1,x_2,x_3;\epsilon ,\epsilon _1,\epsilon _2,\epsilon _3).`$ (13) As seen from Eqs. (10) and (13) the binding energy can be written as $`_b={\displaystyle 𝑑x𝑑x^{}\underset{c}{}\frac{d\epsilon }{2\pi i}\left(tr\widehat{k}_x\delta (xx^{})+M(x,x^{};\epsilon )\right)G(x^{},x;\epsilon )}`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x𝑑x_1\underset{c}{}\frac{d\epsilon d\epsilon _1}{(2\pi i)^2}f_2(|𝐫𝐫_1|;\omega )K_2(x,x_1;x,x_1;\epsilon ,\epsilon _1)}`$ $``$ $`{\displaystyle \frac{2}{3}}{\displaystyle 𝑑x𝑑x_1𝑑x_2\underset{c}{}\frac{d\epsilon d\epsilon _1d\epsilon _2}{(2\pi i)^3}f_3(|𝐫𝐫_1|,|𝐫𝐫_2|;\omega ,\omega _1)K_3(x,x_1,x_2;x,x_1,x_2;\epsilon _1,\epsilon _2)}`$ $``$ $`{\displaystyle \frac{3}{4}}{\displaystyle }{\displaystyle }{\displaystyle }{\displaystyle }dxdx_1dx_2dx_3{\displaystyle }{\displaystyle \underset{c}{}}{\displaystyle }{\displaystyle }{\displaystyle \frac{d\epsilon d\epsilon _1d\epsilon _2d\epsilon _3}{(2\pi i)^4}}f_4(|𝐫𝐫_1|,|𝐫𝐫_2|,|𝐫𝐫_3|,\omega ,\omega _1,\omega _2)\times `$ $`\times `$ $`K_4(x,x_1,x_2,x_3;x,x_1,x_2,x_3;\epsilon ,\epsilon _1,\epsilon _2,\epsilon _3).`$ (14) Taking into account the spectral representation of $`G(x,x^{};\epsilon )`$, see Eq.(5) in Ref., and performing the integration over $`\epsilon `$ we get the following expression for the first term in the rhs: $$_b^{(1)}=\underset{j}{\overset{(A1)}{}}𝑑x\mathrm{\Psi }_j^+(x)\left(\widehat{k}_x\mathrm{\Psi }_j(x)+M(x,x^{};E_j)\mathrm{\Psi }_j(x^{})𝑑x^{}\right).$$ (15) As follows from the Dyson equation (12) and the above spectral representation the amplitudes $`\mathrm{\Psi }_j(x)`$ obey the equation $$\widehat{k}_x\mathrm{\Psi }_j(x)+M(x,x^{};E_j)\mathrm{\Psi }_j(x^{})𝑑x^{}=E_j\mathrm{\Psi }_j(x)$$ (16) and therefore $$_b^{(1)}=\underset{j}{\overset{(A1)}{}}E_j|\mathrm{\Psi }_j(x)|^2𝑑x=\underset{\lambda }{}\underset{j}{\overset{(A1)}{}}E_js_j^{(\lambda )}$$ (17) (we accounted for the expansion (3)). In general case the many-particle Green functions $`K_n(x_1,\mathrm{}x_n;x_1^{},\mathrm{}x_n^{};\epsilon _1,\mathrm{}\epsilon _n)`$ obey the infinite system of integro-differential equations the Eq.(11) being the first one. As mentioned above the exact methods of solution do not exist. It should be also mentioned that the approximate methods, see and the references therein, are developed for the instantaneous two-particle forces only. For these reasons the exact calculation of nuclear binding energy is impossible at present. Therefore the calculation of the static energy is of importance. Both the static nuclear field and the static energy are obtained from the above relations by putting $$K_n(x_1,\mathrm{}x_n;x_1^{},\mathrm{}x_n^{};\epsilon _1,\mathrm{}\epsilon _n)=\underset{i=1}{\overset{n}{}}G(x_i,x_i^{};\epsilon _i),$$ (18) i.e. neglecting the difference between the many-particle Green functions and the unsymmetrized products of the single-particle ones. This difference arises from both the antisymmetrization and the higher-order terms of the perturbation theory, i.e. just the effects leading to the proper correlation energy. Putting Eq.(18) into Eqs.(13) and (14) we get (all the energy transfers $`\omega _i`$ vanish in this case) $`U_{st}(r)={\displaystyle f_2(|𝐫𝐫_1|)\rho (r_1)d^3𝐫_1}+{\displaystyle f_3(|𝐫𝐫_1|,|𝐫𝐫_2|)\rho (r_1)\rho (r_2)d^3𝐫_1d^3𝐫_2}+`$ $`+`$ $`{\displaystyle f_4(|𝐫𝐫_1|,|𝐫𝐫_2|,|𝐫𝐫_3|)\rho (r_1)\rho (r_2)\rho (r_3)d^3𝐫_1d^3𝐫_2d^3𝐫_3};`$ $`_{st}={\displaystyle \underset{j}{\overset{(A1)}{}}}{\displaystyle \mathrm{\Psi }_j^+(x)\left(\widehat{k}_x+U_{st}(r)\right)\mathrm{\Psi }_j(x)𝑑x}{\displaystyle \frac{1}{2}}{\displaystyle f_2(|𝐫𝐫_1|)\rho (r)\rho (r_1)d^3𝐫d^3𝐫_1}`$ $``$ $`{\displaystyle \frac{2}{3}}{\displaystyle f_3(|𝐫𝐫_1|,|𝐫𝐫_2|)\rho (r)\rho (r_1)\rho (r_2)d^3𝐫d^3𝐫_1d^3𝐫_2}`$ (20) $``$ $`{\displaystyle \frac{3}{4}}{\displaystyle f_4(|𝐫𝐫_1|,|𝐫𝐫_2|,|𝐫𝐫_3|)\rho (r)\rho (r_1)\rho (r_2)\rho (r_3)d^3𝐫d^3𝐫_1d^3𝐫_2d^3𝐫_3}.`$ As follows from the relations (3) and (5) the first term of the rhs is $$_{st}^{(1)}=\underset{\lambda }{}n_\lambda \epsilon _\lambda .$$ (21) But as seen from Eq.(17) and the sum rule $$\epsilon _\lambda =\underset{j}{\overset{(A1)}{}}E_js_j^{(\lambda )}+\underset{k}{\overset{(A+1)}{}}E_ks_k^{(\lambda )}$$ (22) for the doorway state energies, see Eq.(14c) of Ref., the quantity $`_{st}^{(1)}`$ contains unphysical contributions from the energies of the $`(A+1)`$ nuclear states. To eliminate them let us divide $`_{st}^{(1)}`$ into the two parts $$_{st}^{(1)}=_{st}^<+_{st}^>,_{st}^<=\underset{\lambda F}{}n_\lambda \epsilon _\lambda ,_{st}^>=\underset{\lambda >F}{}n_\lambda \epsilon _\lambda ,$$ (23) including summations over the states with $`\epsilon _\lambda \epsilon _F`$ and $`\epsilon _\lambda >\epsilon _F`$ respectively. The doorway states entering $`_{st}^<`$ are mainly distributed over the states of the $`(A1)`$ nucleus the first term of Eq.(22) thus giving the dominant contribution to the $`\epsilon _\lambda `$ values in this case. The small unphysical contribution from the second term of (22) arises from the states of the $`(A+1)`$ nucleus with the same quantum numbers as those of the low-lying states of the $`(A1)`$ nucleus. Such states lie either in the continuum or near its border. Therefore the reasonable estimate for their energies is $`E_k0`$ the unphysical contribution thus being negligible for this case. The situation is opposite for $`_{st}^>`$ because the doorway states with $`\epsilon _\lambda >\epsilon _F`$ are mainly distributed over the states of the $`(A+1)`$ nucleus, the dominant contribution to the $`\epsilon _\lambda `$ values thus being unphysical from the viewpoint of the binding energy. In such conditions it is reasonable to use a more appropriate relation for $`_{st}^>`$, $$\stackrel{~}{}_{st}^>=\underset{\lambda >F}{}\underset{j}{\overset{(A1)}{}}E_js_j^{(\lambda )}.$$ (24) So the $`(A1)`$ nuclear states with the same quantum numbers as those of the low-lying states of the $`(A+1)`$ nucleus contribute in this case. There are no such states among the low-lying ones of the $`(A1)`$ nucleus, and therefore $$E_j<E_F=_0(A)_g(A1),$$ (25) see Ref. for the details. Nevertheless it is reasonable to use the estimate $$E_j=E_F,$$ (26) providing the least absolute value for the $`\stackrel{~}{}_{st}^>`$ quantity. In this way we get $$\stackrel{~}{}_{st}^>=E_F\underset{\lambda >F}{}\underset{j}{\overset{(A1)}{}}s_j^{(\lambda )}=E_F\underset{\lambda >F}{}n_\lambda =E_F\stackrel{~}{N}_{out}.$$ (27) Accounting for the fact that the many-particle forces are introduced as the contact ones in the model-independent approach of Ref., we finally get (see Ref. for the notations) $`_{st}={\displaystyle \underset{i=n}{\overset{p}{}}}\left({\displaystyle \underset{\lambda F}{}}n_\lambda \epsilon _\lambda +\stackrel{~}{N}_{out}E_F\right)_i`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle \left(S(r)\rho _s(r)+V_\omega (r)\rho (r)+S^{}(r)\rho _s^{}(r)+V_\rho (r)\stackrel{~}{\rho }^{}(r)+C(r)\rho _{ch}(r)\right)d^3𝐫}`$ $``$ $`{\displaystyle }({\displaystyle \frac{2}{3}}a_3\rho ^3(r)+{\displaystyle \frac{3}{4}}a_4\rho ^4(r)+{\displaystyle \frac{1}{2}}[a_3^{}\rho (r)+a_4^{}\rho ^2(r)](\rho ^{}(r))^2]d^3𝐫.`$ (28) We used the following ansatz for the occupation numbers $$n_\lambda =\frac{1}{2}\left[1\frac{\epsilon _\lambda \mu }{\sqrt{(\epsilon _\lambda \mu )^2+C^2}}\right],\underset{\lambda }{}n_\lambda =N,\underset{\lambda >F}{}n_\lambda =\stackrel{~}{N}_{out},$$ (29) putting $`\stackrel{~}{N}_{out}=N_{out}`$, see Sect.2. This is incorrect because the sum $`_{\lambda >F}n_\lambda `$ is different from $`N_{out}`$, Eq.(9). Indeed, both diagonal and nondiagonal elements of the density matrix, Eq.(4), contribute to the $`N_{out}`$ values. This shortcoming will be corrected in future by calculating the $`n_\lambda `$ and $`\rho _{\lambda \nu }`$ values, see the discussion in Sect.2 (of course, the ansatz (29) will become unnecessary). The results of the calculations are shown in Table 2. As seen from the table the static energy is the sum of a number of contributions with different sign, the dominant ones greatly exceeding the total static energy as well as the observed binding one. This is the source of the ambiguities because all disregarded effects may be of importance in such a situation. One of such effects is the finite range of the many-particle forces. The possible way of the estimation is illustrated for the three-particle forces. As seen from Eq.(10) the original contribution is $`_3={\displaystyle \frac{1}{3}}{\displaystyle f_3(|𝐫𝐫_1|,|𝐫𝐫_2|)\rho (r)\rho (r_1)\rho (r_2)d^3𝐫d^3𝐫_1d^3𝐫_2}=`$ $`=`$ $`{\displaystyle \frac{1}{3}}{\displaystyle \rho (r)f_3(\xi ,\eta )\rho (|𝐫+𝝃|)\rho (|𝐫+𝜼|)d^3𝐫d^3𝝃d^3𝜼}`$ $``$ $`{\displaystyle \frac{1}{3}}{\displaystyle }{\displaystyle }{\displaystyle }\rho (r)f_3(\xi ,\eta )[\rho (r)+𝝃\mathbf{}\rho (r)+{\displaystyle \frac{1}{2}}\xi _1\xi _k_i_k\rho (r)+\mathrm{}]\times `$ $`\times `$ $`\left[\rho (r)+𝜼\mathbf{}\rho (r)+{\displaystyle \frac{1}{2}}\eta _j\eta _{\mathrm{}}_j_{\mathrm{}}\rho (r)+\mathrm{}\right]d^3𝐫d^3𝝃d^3𝜼=`$ (30) $`=`$ $`{\displaystyle \frac{1}{3}}{\displaystyle \rho (r)f_3(\xi ,\eta )\left[\rho (r)+\frac{1}{6}\xi ^2\mathrm{\Delta }\rho (r)+\mathrm{}\right]\left[\rho (r)+\frac{1}{6}\eta ^2\mathrm{\Delta }\rho (r)+\mathrm{}\right]d^3𝐫d^3𝝃d^3𝜼}=`$ $`=`$ $`{\displaystyle \frac{1}{3}}a_3{\displaystyle \rho (r)[\rho (r)+r_0^2\mathrm{\Delta }\rho (r)+\mathrm{}]^2d^3𝐫}{\displaystyle \frac{1}{3}}a_3{\displaystyle \rho ^3(r)d^3𝐫}+{\displaystyle \frac{2}{3}}a_3{\displaystyle \rho ^2(r)\mathrm{\Delta }\rho (r)d^3𝐫},`$ where (see Eq.(37) of Ref.) $$a_3=f_3(\xi ,\eta )d^3𝝃d^3𝜼,a_3r_0^2=\frac{1}{6}\xi ^2f_3(\xi ,\eta )d^3𝝃d^3𝜼=\frac{1}{6}\eta ^2f_3(\xi ,\eta )d^3𝝃d^3𝜼.$$ (31) In this way we get $`\delta _{r_0^2}_{st}=r_0^2{\displaystyle \left(\frac{2}{3}a_3\rho ^2(r)+\frac{3}{4}a_4\rho ^3(r)\right)\mathrm{\Delta }\rho (r)d^3𝐫}=Dr_0^2`$ $`D={\displaystyle \left(\frac{4}{3}a_3\rho (r)+\frac{9}{4}a_4\rho ^2(r)\right)(\rho ^{}(r))^2d^3𝐫}.`$ (32) The calculated $`D`$ values are shown in Table 3. The reasonable value of the range parameter is the $`\omega `$ meson Compton wavelength, i.e. a typical scale of the strong interaction. So $`r_0^2=m_\omega ^2=0.1`$fm<sup>2</sup>, the effect thus being negligibly small. The second source of the possible ambiguity is seen from the results of Ref. for the contributions of two-particle, three-particle and four-particle forces to the nuclear static field which are shown in Table 4, see Eqs. (44) and (45) of Ref.. As follows from the table the convergence of this sequence is not seen. This observation suggests the possible existence of the contributions from higher many-particle forces, and therefore the whole sequence should be summed up. But this can be done only within some reasonable model for the many-particle forces including all higher-order ones as well as the finite range and the mechanism for the saturation of nuclear density. The corresponding investigation is in progress.
warning/0003/astro-ph0003064.html
ar5iv
text
# One relation for self-gravitating bodies ## I Introduction For homogeneous triaxial ellipsoid with semiaxes $`A,B,C`$ and density $`\rho `$ the next relations are valid (Landau and Lifshits 1975).The gravitational potential at the inner point ($`X,Y,Z`$), $`AXA`$, $`BYB`$, $`CZC`$, is $$\begin{array}{c}U_{ell}(X,Y,Z)=\pi \rho GABC_0^{\mathrm{}}\left(1\frac{X^2}{A^2+s}\frac{Y^2}{B^2+s}\frac{Z^2}{C^2+s}\right)\frac{ds}{Q_s};\hfill \\ Q_s=\sqrt{(A^2+s)(B^2+s)(C^2+s)}.\hfill \end{array}$$ (1) The potential energy of the homogeneous triaxial ellipsoid is: $$W_{ell}=\frac{3}{10}GM^2_0^{\mathrm{}}\frac{ds}{Q_s}.$$ (2) Here $`M_{ell}=4/3\pi \rho ABC`$ is the ellipsoid’s mass and $`G`$ stands for Newtonian constant of gravitation. Note that gravitational energy of self-gravitating body is of negative sign but we loosely write all $`W`$s with positive sign. In general case $`ABC`$ the integrals in (1) and (2) are expressed only in terms of the incomplete elliptic integrals and this precludes any detailed analysis. However, if we consider only potential at the center of ellipsoid, $`U_{ell}(0,0,0)`$, then we get the remarkable relation: $$WUM=\frac{W_{ell}}{U_{ell}(0,0,0)M_{ell}}=\frac{2}{5},$$ (3) valid for any values of semi-axes. We shortly refer to this relation (3) as $`WUM`$-ratio. Recently Seidov and Skvirsky (2000a) presented the gravitational potential and the potential energy for the homogeneous rectangular parallelepiped (hereafter RP) which allows to analyse this $`WUM`$-ratio for the new class of the homogeneous self-gravitating bodies. In this paper, we show in sections II and III that the value of $`WUM`$ for RPs has minimal value $`.395437`$ (see Eq. (10)) for the cube (all three dimensions of RP equal to each other), tends to 1/2 for one dimension of RP far larger larger than two others (long thin ”stick” with square cross-section), and tends to $`\frac{1}{2}\frac{\sqrt{2}1}{6\mathrm{ln}(\sqrt{2}+1)}=.421673`$ as one dimension is far less than two others (thin square ”plate”), see Fig. 1. Also, in section IV we discuss the $`WUM`$-ratio for homogeneous gravitating bodies studied recently by Kondrat’ev and Antonov (1993). We show that values of $`WUM`$-ratio for the homogeneous summetrical lenses are in the interval from $`128/105\pi `$, for infinitesimally thin symmetrical lens, to $`17/20`$, for two homogeneous equal spheres just touching each other. In section VI we analyse $`WUM`$-ratio for spherical polytropic stars, and show that $`WUM`$-ratio varies from $`2/5`$ to $`3/32\pi `$ for polytropic index $`n`$ varying from $`0`$ to $`5`$. In the section VII we discuss the interesting class of two-phase spheres and show that unlike the polytropes, in this case the $`WUM`$-ratio’s interval is larger: it is possible to get very small values of $`WUM`$ if the ratio of two densities $`q=\rho _2/\rho _1`$ is large enough and if the relative value of core’s radius is not too small. At last, in the sections VIII and IX we consider another two simple classes of non-homogeneous bodies both allowing analytical treatment. ## II Potential at the center of RP Using results by Seidov and Skvirsky (2000a) we write down the gravitational potential at the center of the homogeneous RP with density $`\rho `$ and with edge lengths $`2a,\mathrm{\hspace{0.17em}2}b,\mathrm{\hspace{0.17em}2}c`$: $$\begin{array}{c}U_{RP}(0,0,0)=4G\rho [ab\mathrm{ln}\frac{d+c}{dc}+bc\mathrm{ln}\frac{d+a}{da}+cd\mathrm{ln}\frac{d+b}{db}\hfill \\ \\ a^2\mathrm{arctan}\frac{bc}{ad}b^2\mathrm{arctan}\frac{ac}{bd}c^2\mathrm{arctan}\frac{ab}{cd}].\hfill \end{array}$$ (4) Here $`d=(a^2+b^2+c^2)^{1/2}`$ is the main diagonal of RP. Three particular cases are of the larger interest: a)cube corresponding to case $`c=b=a`$, $$U_{cube}(0,0,0)=G\rho a^2\left[24\mathrm{ln}\frac{1+\sqrt{3}}{\sqrt{2}}2\pi \right]=9.52017G\rho a^2;$$ (5) b)long thin stick with square cross-section corresponding to case $`a>>b=c`$: $$U_{stick}(0,0,0)=G\rho b^2(8\mathrm{ln}\frac{b}{a}+122\pi +4\mathrm{ln}2);$$ (6) c)thin square plate corresponding to the case $`a<<b=c`$: $$U_{plate}(0,0,0)=G\rho [16ab\mathrm{ln}(\sqrt{2}+1)2\pi a^2].$$ (7) ## III Potential energy of RP According to Seidov and Skvirsky (2000a) the gravitational potential energy of the homogeneous rectangular parallelepiped is equal to: $$\begin{array}{c}WRP=G\rho ^2[f(a,b,c)+f(b,c,a)+f(c,a,b)];f(a,b,c)=c_5a^5+c_4a^4+c_3a^3+c_2a^2;\hfill \\ \\ c_5=\frac{32}{15};c_4=\frac{32}{15}(dd1d3)\frac{16b}{3}\mathrm{ln}\frac{(d1b)(d+b)}{ad3}\frac{16c}{3}\mathrm{ln}\frac{(d3c)(d+c)}{ad1};\hfill \\ \\ c_3=\frac{64bc}{3}\mathrm{arctan}\frac{bc}{ad};d=\sqrt{a^2+b^2+c2};d1=\sqrt{a^2+b^2};d3=\sqrt{a^2+c^2};\hfill \\ \\ c_2=\frac{32b^2}{5}(d1d)+\frac{32c^2}{5}(d3d)16bc^2\mathrm{ln}\frac{db}{d+b}16b^2c\mathrm{ln}\frac{dc}{d+c}.\hfill \end{array}$$ (8) ### A Potential energy and $`WUM`$-ratio of cube From (8), taking $`c=b=a`$, we get the potential energy of homogeneous cube with edge length $`2a`$: $$\begin{array}{c}W_{cube}=32G\rho ^2a^5\{\frac{2\sqrt{3}\sqrt{2}1}{5}+\frac{\pi }{3}+\mathrm{ln}[(\sqrt{2}1)(2\sqrt{3})]\}=30.117G\rho ^2a^5.\hfill \end{array}$$ (9) From (5) and (9) we get $`WUM`$-ratio for homogeneous cube: $$WUM_{cube}=2\frac{\frac{2\sqrt{3}\sqrt{2}1}{5}+\frac{\pi }{3}+\mathrm{ln}[(\sqrt{2}1)(2\sqrt{3})]}{24\mathrm{ln}\frac{1+\sqrt{3}}{\sqrt{2}}2\pi }=.395437.$$ (10) ### B Potential energy and $`WUM`$-ratio of thin long stick We take $`a>>b=c`$ that corresponds to the case of the thin long stick with the square cross-section. Leading term in expansion of WRP (8) gives the potential energy of the thin long stick: $$W_{stick}=\frac{32}{3}G\rho ^2ab^4\mathrm{ln}\frac{a}{b}.$$ (11) From this and (6) we get for stick: $$WUM=\frac{W_{stick}}{8ab^2U_{stick}(0,0,0)}=\frac{1}{2}.$$ (12) ### C Potential energy and $`WUM`$-ratio of thin square plate Taking one of RP’s dimension infinitesimally small, $`a0`$, we get, from Eq.(8), the potential energy of the thin rectangular plate. If we additionally take $`b=c`$, then we get the potential energy of the thin square plate ($`a<<b`$): $$W_{pl}=64G\rho ^2b^3a^2\left(\mathrm{ln}(\sqrt{2}+1)\frac{\sqrt{2}1}{3}\right)=47.5714G\rho ^2b^3a^2.$$ (13) From (7) and (13) we have another limit for value of $`WUM`$: $$\frac{WS}{8ab^2U_{pl}(0,0,0)}=\frac{1}{2}\frac{\sqrt{2}1}{6\mathrm{ln}(\sqrt{2}+1)}=.421673.$$ (14) ### D Relation between potential energy, gravitational potential and mass of RP General behavior of relation between the potential energy, the gravitational potential at the center, and mass of the homogeneous RP with two equal edge-lengths is shown in the Fig. 1, which is a result of numerical calculation by the formulas (4) and (8). For homogeneous ellipsoid $`WUM=2/5`$, see (3), solid line in Fig. 1. ## IV Homogeneous symmetric lenses Recently Kondrat’ev and Antonov (1993) (hereafter KA) have obtained the analytical formulas for the gravitational potential and the gravitational energy of some axial-symmetric figures, namely homogeneous lenses with spherical surfaces of different radii. In forthcoming paper (Seidov and Skvirsky 2000b) we present some new solutions for homogeneous bodies of revolution. Here we present the review of $`WUM`$ for most suitable kind of those bodies, discussed by KA, namely the homogeneous symmetrical lenses. A segment of sphere, or a planoconvex lens, is obtained by cutting a sphere with a plane. If $`R`$ is a radius of a sphere, $`h`$ is height of segment, and $`2a`$ is a radius of segment’s base, then $`a=\sqrt{2hRh^2}`$. A symmetric homogeneous lens (SL) is obtained by placing together the bases of two identical segments of sphere. According to KA we have for the gravitational potential at the center of such SL: $$U_{SL}(0)=\frac{4\pi G\rho }{3}\left(hR+R^2\frac{h^2}{2}+\frac{R^3a^3}{hR}\right).$$ (15) The gravitational energy of SL is: $$\begin{array}{c}W_{SL}=\frac{4\pi G\rho ^2}{9}[10R^4a\frac{16}{3}R^2a^3\frac{8}{3}a^5+\pi h^4(\frac{2}{5}h2R\frac{R^2}{Rh})+2R^4(\frac{R^2}{Rh}6(Rh))\mathrm{arctan}\frac{a}{Rh})].\hfill \end{array}$$ (16) And the total mass of the homogeneous SL is: $$M_{SL}=\frac{2\pi \rho h^2}{3}(3Rh).$$ (17) We may consider $`WUM`$-ratio for such bodies as $`W_{SL}/U_{SL}M_{SL}`$. Defined so, $`WUM`$ for the homogeneous symmetric lenses has general dependence on parameter $`h/R`$, according the formulas (15), (16) and (17), as shown in the Fig. 2. Note that in the limit $`h/R0`$ we have infinitesimally thin symmetrical lens with radius of curvature tending to infinity, and we have: $$Limit[WUM_{SL}]_{h/R>0}=\frac{128}{105\pi }=.388035.$$ (18) Interestingly, this infinitesimally thin round lens does not coincide at all with the case of the infinitesimally thin quadratic plane which has the much more larger value of $`WUM=.421673`$ (see Eq.(14)). In another limit $`h/R1`$ we have a full homogeneous sphere, and evidently: $$Limit[WUM_{SL}]_{h/R>1}=2/5.$$ (19) There is still another $`analytical`$ case at $`h2R`$ when we have two homogeneous spheres touching each other so that distance between centers of spheres is $`d=2R;`$ then ”potential at the center of SL” is $`U_{SL}(0)=2\frac{GM}{R}`$; potential energy is sum of two terms: proper potential energy of each spheres $`2\frac{3}{5}\frac{GM^2}{R}`$, and $`interaction`$ energy, $`\frac{GM^2}{d}`$ or $`\frac{GM^2}{2R}`$; we have: $$WUM_{SL}|_{h=2R}=(\frac{1}{2}+\frac{6}{5})/2=\frac{17}{20}.$$ ## V Heterogeneous spherical bodies Now we consider the problem of $`WUM`$-ratio from another point of view. Abovementioned homogeneous gravitating bodies (ellipsoids, right parallelepipeds, and symmetrical spherical lenses) differ from each other only by their forms and so $`WUM`$-ratio may be referred to as form-factor. As it is evident, $`WUM`$-ratio should be also function of density distribution over the body. If we confine ourselves by spherically-symmetric distribution of density $`\rho (r)`$, then we have general expressions for the gravitational potential at radius $`r`$: $$U(r)=\frac{Gm(r)}{r}+_r^R\mathrm{\hspace{0.17em}4}\pi G\rho (r)r𝑑r;$$ (20) central potential: $$U(0)=_0^R\mathrm{\hspace{0.17em}4}\pi G\rho (r)r𝑑r;$$ (21) potential energy: $$W=\frac{1}{2}_0^MU(r)𝑑m(r);$$ (22) and mass: $$dm=4\pi \rho (r)r^2dr;M=_0^R\mathrm{\hspace{0.17em}4}\pi \rho (r)r^2𝑑r.$$ (23) From these expressions we write down $`WUM`$ for spherical-summetrical heterogeneous bodies as: $$WUM=\frac{W}{U(0)M}=\frac{1}{2}\frac{\overline{U}}{U(0)};\overline{U}=\frac{1}{M}_0^MU(m)𝑑m.$$ (24) As a result, $`WUM`$-ratio of spherically-symmetric bodies is reduced to the ratio of the mean value of monotonic function, $`U(m)`$, to its particular value, $`U(0)`$, (ratio being additionally divided by 2). The boundary values of this ratio can be found pure mathematically for any given class of functions $`\rho (r)`$. We will not deal with this abstract (though interesting) problem; instead, in the next sections we consider two cases of more or less realistic bodies, namely polytropes and two-phase spheres. ## VI Polytropes One case of heterogeneous bodies which apparently should be considered first is the case of classical polytropic stars. Using formulas from Chandrasekar’s (1957) classical text we have (some of these formulas are valid not only for polytropic stars but we do not stop on these details) the next relations. a)central potential is expressed via other parameters of star as follows (Chan 100/85= Chandrasekhar (1957), p.100, Eq. (85)): $$U_p(0)=(n+1)\frac{P_c}{\rho _c}+\frac{GM}{R}.$$ (25) Here $`n`$ is the polytropic index, $`P_c`$ and $`\rho _c`$ are central values of pressure and density, $`M`$ and $`R`$ are the mass and radius of the star. The next formulas include the parameters of the Lane-Emden function (LEF) at the first zero point: $$\begin{array}{c}\xi =\xi _1,\theta (\xi =\xi _1)\theta _1=0;[\frac{d\theta (\xi )}{d\xi }]_{\xi =\xi _1}\theta _1^{}<\mathrm{\hspace{0.17em}0};\mu _1=\xi _1^2\theta _1^{}.\hfill \end{array}$$ (26) Central pressure is (Chan 99/80,81): $$P_c=\frac{1}{4\pi (n+1)(\theta _{1}^{}{}_{}{}^{})^2}\frac{GM^2}{R^4}.$$ (27) Central density $`\rho _c`$ is related with the mean density $`\overline{\rho }`$ of the star as follows (Chan 78/99): $$\rho _c=\frac{1}{3}\frac{\xi }{\theta _{1}^{}{}_{}{}^{}}\overline{\rho }.$$ (28) Combining these formulas we get the final expression for the central potential of polytropic star: $$U_p(0)=\left(1+\frac{1}{\xi _1\theta _{1}^{}{}_{}{}^{}}\right)\frac{GM}{R}.$$ (29) b)potential energy We have for polyropic star the famous formula (Chan 101/90): $$W_p=\frac{3}{5n}\frac{GM^2}{R}.$$ (30) c)WUM-ratio: $$WUM_p=\frac{3}{5n}\frac{1}{1+\frac{\xi _1}{\mu _1}}.$$ (31) ### A $`WUM`$-ratio for polytropes Using values of parameters of polytropic stars given in (Chan, Table 4, p. 96) we calculated the values of $`WUM`$-ratio for values of polytropic index $`n`$ from $`0`$ to $`4.9`$, see Fig. 3. However the limit $`n5`$ and in general region of values of $`n`$ close to 5 should be considered separately. First we note that at $`n5`$, $`\xi _1\mathrm{}`$ and there is indeterminacy of $`\frac{\mathrm{}}{\mathrm{}}`$ kind in formula for $`WUM_p`$ Eq. (31). To solve this problem we use results of Seidov and Kuzakhmedov (1979) who obtained, in particular, the dependence of $`\xi _1`$ for $`n`$ close to 5: $$(0<(5n)/5<<\mathrm{\hspace{0.17em}1}),\xi _1=\frac{32\sqrt{3}}{\pi }\frac{1}{5n}.$$ (32) Also it was shown by Seidov and Kuzakhmedov (1979) that for values of $`n`$ close to 5, the parameter $`\mu _1`$ is an increasing function of $`n`$: $$\mu _1=\sqrt{3}(1+\frac{1}{12}\frac{1}{5n}),$$ (33) that means that $`\mu _1`$ has minimum at $`n`$ about 4.82. This may or may not lead to minimum of $`WUM_p`$ as function of $`n`$. At point $`n=5`$ using LEF of index 5: $$n=5,\theta (\xi )=(1+\frac{1}{3}\xi ^2)^{1/2},$$ (34) we get $$n=5,WUM_p=\frac{3}{32}\pi =.294524.$$ (35) Additionally we have two other analytical results for $`WUM`$ of polytropes : at $`n=0`$, $`WUM=2/5`$ and at $`n=1`$, $`WUM=3/8`$. We recalculated the parameters of polytropes in the interval of $`n`$ from 4 to 5 and found no minimum for $`WUM`$, see Fig. 3. However we note that our boundary values differ from ones calculated by Jabbar (1993) in the sense that ours are less than his. As one example, at $`n=4.7`$ Jabbar gives $`\xi _1=54.810686`$ as zero of LEF, while our calculations using Mathematica’s command NDSolve give $`\theta (54.810686)=4.5910^7`$ and our value of $`\xi _1`$ is in interval between $`\xi =54.8098`$ (at this point $`\theta =4.9810^{}8`$) and $`\xi =54.8099`$ (at this point $`\theta <0`$). In general our values of $`\xi _1`$ are less than Jabbar’s. ## VII Two-phase sphere If $`\rho _1`$ and $`\rho _2=q\rho _1`$ are densities in envelope and core of sphere, and $`R`$ and $`r=xR`$ are total radius of sphere and radius of core, then we have next relations: a)gravitational potential at the center: $$U_{2ph}(0)=2\pi G\rho _1R^2[1+(q1)x^2];$$ (36) b)potential energy: $$W_{2ph}=\frac{16\pi ^2G}{15}\rho _{1}^{}{}_{}{}^{2}R^5[1+\frac{5}{2}(q1)x^3+(q1)(q\frac{3}{2})x^5];$$ (37) c)total mass: $$M_{2ph}=\frac{4\pi }{3}\rho _1R^3[1+(q1)x^3];$$ (38) d)$`WUMratio`$: $$\begin{array}{c}WUM_{2ph}=\frac{2}{5}\frac{1+\frac{5}{2}(q1)x^3+(q1)(q\frac{3}{2})x^5}{[1+(q1)x^2][1+(q1)x^3]}.\hfill \end{array}$$ (39) In Fig.4 we present dependence of $`WUM`$-ratio, for two-phase spheres with various values of $`q`$, as function of relative radius of core $`x=r/R`$. Curves from uppper one to lower one correspond to values of $`q=3,\mathrm{\hspace{0.17em}5},\mathrm{\hspace{0.17em}10},`$ and $`20`$, respectively. Both values of minimuma of $`WUM`$ and their ”places”, corresponding values of $`x`$, are decreasing functions of $`q`$. For values of $`x`$ corresponding to minimuma of $`WUM`$ the next simple equation is valid: $$6q^2x^5+5q(1x)x^3(1+2x)(1x)^3(4+7x+4x^2)=0.$$ (40) or $$q=\frac{(1x)\left(\sqrt{96/x+97+28x+4x^2}510\sqrt{x}\right)}{12x^2}.$$ (41) We used these formulas to calculate the dash line in Fig. 4. ## VIII Stepenars Here we briefly consider the case of simple spherically-symmetric density distribution law allowing analytical expression for WUM. We take: $$\rho (r)=\rho _c(1r/R)^\nu ,$$ (42) where $`\rho _c`$ is central density and $`\nu `$ is a free parameter. By historical reasons we refer to the gravitating bodies with density distribution (42) as ”stepenars”, (Seidov, Kasumov, and Guseinov 1971). We have: mass: $$M=\frac{8\pi \rho _cR^3}{(1+\nu )(2+\nu )(3+\nu )};$$ (43) mean-to-central density ratio: $$\frac{\overline{\rho }}{\rho _c}=\frac{6}{(1+\nu )(2+\nu )(3+\nu )};$$ (44) central potential: $$U(0)=\frac{4\pi G\rho _cR^2}{(1+\nu )(2+\nu )};$$ (45) potential energy: $$W=\frac{8\pi ^2G\rho _cR^2(8+5\nu )}{(1+\nu )^2(2+\nu )^2(3+2\nu )(5+2\nu )};$$ (46) WUM-ratio: $$WUM_\nu =\frac{(3+\nu )(8+5\nu )}{4(3+2\nu )(5+2\nu )}.$$ (47) In Fig. 5 the dependence of $`WUM`$-ratio for stepenars as function of parameter $`\nu `$ is shown. In spite of large variation of matter concentration from $`\rho =const`$ at $`\nu =0`$, to $`\overline{\rho }/\rho _c0`$ at $`\nu \mathrm{}`$, $`WUM`$-ratio again as in polytrope’s case lies in rather limited interval from $`2/5`$ to $`5/16`$. ## IX Alphars Here we consider another example of ”exotic” but simple density distribution, Seidov and Seidova (1971): $$\rho (r)=\rho _c(r/R)^\alpha ,0\alpha 2.$$ (48) We have: $$WUM_\alpha =\frac{2\alpha }{52\alpha }.$$ (49) ## X Discussion There is a rather classic problem of looking for general theorems of stellar structure, see e.g. chapter 3 in the classic text Chandrasekhar (1957). The problem considered in this paper may be also referred to as that dealing with general structure of celestial self-gravitating bodies. We start from interesting observation on one constant ratio, namely, (potential energy W)/((central potential U ) x (total mass M)), in homogeneous ellipsoids and then try to look for behavior of this ratio for another homogeneous bodies: rectangular parallelelepipeds and symmetrical lenses. We found that in both cases $`WUM`$-ratios are confined in rather narrow interval. Suprisingly, dependence of $`WUM`$ for homogeneous rectangular parallelepipeds (RP) on edge lengths ratio is non-monotonic: it has minimal value $`.395437`$ for cube while any deviation from cube form to prolate RP (one dimension being smaller than two others) or elongated RP (one dimension being larger that two others) leads to the increase of value of $`WUM`$. In this respect the behavior of homogeneous rectangular parallelepipeds is quite unlike the behavior of homogeneous ellipsoids and there is still some mystery even to authors. As to the homogeneous symmetrical lenses (SL) by Kondrat’ev and Antonov (1993), here the dependence of $`WUM`$ on parameters of SL is monotonic, however in this case there is also some suprise in the sense that in the limiting case of thin symmetrical $`spherical`$ lens the $`WUM`$-ratio’s value, ($`128/105\pi `$), differs radically from the case of the infinitesimally thin $`quadratic`$ plate with $`WUM=1/2`$. Then we look for the non-homogeneous however spherically symmetric bodies and found that for the polytropes with polytropic index $`n`$ in the interval $`05`$, $`WUM`$-ratio again lies in narrow interval from $`2/5`$ to $`3/32\pi `$. However for two-phase sphere with large ratio of densities $`q=\rho _2/\rho _1`$ it is possible to get very small values of WUM. The physical reason of it is that if we put in the center of any spherical symmetric star a (very) small but dense spherical body then central potential may be very large while total potential energy of star, being integral value, increases not so drastically. The effect of the strong variation of density in the center of star, the ”first-order phase-transition”, is known since pioneer works of W.H. Ramsey (1950). In last two paragraphs of paper we consider pure mathematical toy models in further attempts to understand the behavior of the $`WUM`$-ratio. We conclude this discussion with notice that central-to-surface potential ratio $`U(0)/U(R)`$ (among other ”global” characteristics of the celestial self-gravitating configurations) is also worth studying. For polytropes, $`U(0)/U(R)=1+\xi _1/\mu _1`$, see section VI. ## Acknowledgements We are grateful to Dr. E. Liverts for valuable discussions.
warning/0003/math0003207.html
ar5iv
text
# Expansiveness of algebraic actions on connected groups ## 1 Introduction Let $`(X,d)`$ be a metric space and $`\rho `$ be a continuous action of a semigroup $`\mathrm{\Gamma }`$ on $`X`$. Then $`(X,\rho )`$ is said to be $`\mathrm{𝑒𝑥𝑝𝑎𝑛𝑠𝑖𝑣𝑒}`$ if there exists an $`ϵ>0`$ such that for any two distinct points $`x,y`$ in $`X`$, $$\underset{\gamma \mathrm{\Gamma }}{sup}d(\rho (\gamma )(x),\rho (\gamma )(y))ϵ.$$ Any such $`ϵ`$ is called an expansive constant of $`(X,\rho )`$. It is easy to check that when $`X`$ is compact, the notion of expansiveness is independent of the metric $`d`$. If $`X`$ is a metrizable topological group and $`\rho `$ is an action of a semigroup $`\mathrm{\Gamma }`$ on $`X`$ by continuous endomorphisms of $`X`$, then $`\rho `$ is said to be expansive if there exists a neighborhood $`U`$ of the identity in $`X`$ such that $$\underset{\gamma }{}\rho (\gamma )^1(U)=\{e\}.$$ Any such neighborhood is called an expansive neighborhood of the identity in $`X`$. An automorphism $`\tau `$ of $`X`$ is said to be expansive if the cyclic group generated by $`\tau `$ acts expansively on $`X`$. It is easy to check that when $`X`$ is compact and $`\rho `$ is an endomorphism action of a semigroup $`\mathrm{\Gamma }`$, these two notions of expansiveness coincide. The notion of expansiveness plays an important role in the study of dynamical systems in general and endomorphism actions in particular. Earlier, expansiveness of automorphisms of connected groups was studied by several authors (cf. , , , and ). A complete description of expansive automorphisms on compact connected groups was obtained in . In particular, it was shown that every compact connected group which admits an expansive automorphism is abelian and finite-dimensional. Also in it was proved that if a compact connected group $`X`$ admits an expansive endomorphism action then it is abelian. In recent years dynamics of endomorphism actions of discrete semigroups on compact abelian groups has been extensively studied using techniques from commutative algebra (cf. ). Among other things, a complete characterization of expansiveness has been obtained when $`X`$ is zero-dimensional ( cf. , Theorem 5.2) or $`\mathrm{\Gamma }=^d`$ for some $`d1`$ (cf. , Theorem 3.9). For further study of expansiveness and it’s relation with other dynamical properties the reader is referred to , and . In this paper we consider expansive endomorphism actions of arbitrary semigroups on connected metrizable topological groups. For any such action $`(G,\rho )`$ we give necessary and sufficient conditions for expansiveness of $`\rho `$, provided $`G`$ is either a Lie group or a compact finite-dimensional group. This paper is organized as follows. In section 2 we prove some elementary results about expansiveness of endomorphism actions on finite-dimensional vector spaces, considered as abelian topological groups under addition. In Section 3 we consider endomorphism actions on connected Lie groups. If $`G`$ is a connected Lie group then by $`L(G)`$ we denote the Lie algebra of $`G`$. If $`\rho `$ is an endomorphism action of a semigroup $`\mathrm{\Gamma }`$ on $`G`$ then by $`\rho _e`$ we denote the induced $`\mathrm{\Gamma }`$-action on $`L(G)`$ defined by $$\rho _e(\gamma )=\mathrm{d}\rho (\gamma )|_e\gamma \mathrm{\Gamma }.$$ We prove the following. Theorem A. Let $`G`$ be a connected Lie group, $`\mathrm{\Gamma }`$ be a semigroup and $`\rho `$ be an endomorphism action of $`\mathrm{\Gamma }`$ on $`G`$. Then $`(G,\rho )`$ is expansive if and only if for all non-zero $`v`$ in $`L(G)`$, the $`\rho _e`$-orbit of $`v`$ is unbounded. In the special case when $`\mathrm{\Gamma }`$ is either abelian or a virtually nilpotent group, applying the above theorem we give a necessary and sufficient condition for expansiveness of $`(G,\rho )`$ in terms of the generalized weights of $`\rho _e`$. As a consequence we prove that if $`\mathrm{\Gamma }`$ is a virtually nilpotent group and $`(G,\rho )`$ is expansive then for some $`\gamma _0`$ in $`\mathrm{\Gamma }`$, $`\rho (\gamma _0)`$ is an expansive automorphism of $`G`$. In section 4 we consider endomorphism actions on compact connected finite-dimensional abelian groups. For any such group $`G`$, by $`\widehat{G}`$ we denote the Pontryagin dual of $`G`$ and by $`L(G)`$ we denote the vector space of all homomorphisms from $`\widehat{G}`$ to $``$ under pointwise addition and scalar multiplication. From the duality theory of compact abelian groups it follows that $`L(G)`$ is finite-dimensional. If $`\rho `$ is an endomorphism action of a semigroup $`\mathrm{\Gamma }`$ on $`G`$ then by $`\rho _e`$ we denote the induced $`\mathrm{\Gamma }`$-action on $`L(G)`$ defined by $$\rho _e(\gamma )(p)(\chi )=p(\chi \rho (\gamma ))pL(G).$$ Also, for such an action $`\rho `$, by $`\widehat{\rho }`$ we denote the induced $`\mathrm{\Gamma }`$-action on $`\widehat{G}`$. We note that $`\widehat{G}`$ can be realised as a module over $`(\mathrm{\Gamma })`$, the group-ring of $`\mathrm{\Gamma }`$, via the action $`\widehat{\rho }`$. We prove the following. Theorem B. Let $`\mathrm{\Gamma }`$ be a semigroup and $`\rho `$ be an endomorphism action of $`\mathrm{\Gamma }`$ on a compact connected finite-dimensional abelian group $`G`$. Then $`(G,\rho )`$ is expansive if and only if the following two conditions are satisfied. a) $`\widehat{G}`$ is finitely generated as a $`[\mathrm{\Gamma }]`$-module. b) For every non-zero $`v`$ in $`L(G)`$, the $`\rho _e`$-orbit of $`v`$ is unbounded. #### Remark 1.1 Since non-abelian compact connected groups do not admit expansive endomorphism actions (cf. ), Theorem B characterizes expansive endomorphism actions on arbitrary compact connected finite-dimensional groups. ## 2 Preliminaries Throughout this paper $`\mathrm{\Gamma }`$ will denote a discrete semigroup i.e. a set with an associative multiplication law. If $`\mathrm{\Gamma }_1,\mathrm{\Gamma }_2`$ are semigroups then a map $`f`$ from $`\mathrm{\Gamma }_1`$ to $`\mathrm{\Gamma }_2`$ is said to be a semigroup homomorphism if for all $`\gamma _1,\gamma _2`$ in $`\mathrm{\Gamma }_1`$, $`f(\gamma _1\gamma _2)=f(\gamma _1)f(\gamma _2)`$. If $`G`$ is a topological group then by $`\mathrm{End}(G)`$ we denote semigroup of all continuous endomorphisms of $`G`$, with composition as the product law. Suppose $`G`$ is a topological group and $`\rho :\mathrm{\Gamma }\mathrm{End}(G)`$ is an endomorphism action of a discrete semigroup $`\mathrm{\Gamma }`$ on $`G`$. Then it is easy to see that $`\mathrm{\Gamma }_1=\rho (\mathrm{\Gamma })\{\mathrm{Id}\}`$ is a subsemigroup of $`\mathrm{End}(G)`$ and $`\rho `$ is expansive if and only if the natural action of $`\mathrm{\Gamma }_1`$ on $`G`$ is expansive. Henceforth we will restrict the discussion of endomorphism actions to the case when $`\mathrm{\Gamma }`$ contains an identity element $`e`$ and $`\rho (e)\mathrm{End}(G)`$ is the identity automorphism. If $`𝐂`$ denotes the semigroup of complex numbers under multiplication then any semigroup homomorphism $`\lambda :\mathrm{\Gamma }𝐂`$ is said to be a $`\mathrm{𝑐ℎ𝑎𝑟𝑎𝑐𝑡𝑒𝑟}`$ of $`\mathrm{\Gamma }`$. If $`\rho :\mathrm{\Gamma }\mathrm{End}(V)`$ is an endomorphism action of $`\mathrm{\Gamma }`$ on a finite-dimensional vector space $`V`$ over $``$, then $`\rho `$ is said to be reducible if there exists characters $`\lambda _1,\mathrm{},\lambda _k`$ and nontrivial $`\rho `$-invariant subspaces $`V_1,\mathrm{},V_kV`$ satisfying the following conditions. 1. $`V=V_1\mathrm{}V_k`$. 2. For $`i=1,\mathrm{},k`$, there exists a basis $`B`$ of $`V_i`$ such that $`\rho (\gamma )|_{V_i}\lambda _i(\gamma )I`$ is strictly upper triangular with respect to $`B`$, for all $`\gamma `$ in $`\mathrm{\Gamma }`$. The characters $`\lambda _1,\mathrm{},\lambda _k`$ are said to be the generalized weights of $`\rho `$ and for all $`i`$, $`V_i`$ is said to be the generalized weight space corresponding to $`\lambda _i`$. #### Example 2.1 If $`\rho (\mathrm{\Gamma })`$ is an abelian subsemigroup $`\mathrm{End}(V)`$ then $`\rho `$ is reducible (see , pp 134). This is also true when $`\rho (\mathrm{\Gamma })`$ is contained in a connected nilpotent subgroup $`H`$ of $`GL(V)`$. This can be seen by looking at the Lie algebra homomorphism $`i:L(H)\text{End}(V)`$ and using certain fundamental facts about representations of nilpotent Lie algebras (see , Proposition 2.4 and Theorem 1.35). The following proposition gives a necessary and sufficient condition for expansiveness of endomorphism actions on a finite-dimensional vector space, considered as an abelian group under addition. ###### Proposition 2.2 Let $`\mathrm{\Gamma }`$ be a semigroup and $`\rho :\mathrm{\Gamma }\mathrm{End}(V)`$ be an endomorphism action of $`\mathrm{\Gamma }`$ on a finite-dimensional vector space $`V`$ over $``$ or $``$. Then we have the following : a) $`\rho `$ is expansive if and only if for all non-zero $`v`$ in $`V`$, the $`\mathrm{\Gamma }`$-orbit of $`v`$ is unbounded. b) If $`V`$ is a real vector space then $`\rho `$ is expansive if and only if the induced endomorphism action of $`\mathrm{\Gamma }`$ on $`V`$ is expansive. #### Proof For any $`v`$ in $`V`$ let $`O_vV`$ denote the $`\rho `$-orbit of $`v`$. Suppose that there is a non-zero vector $`v`$ in $`V`$ with bounded $`\rho `$-orbit. Let $`U`$ be any neighborhood of $`0`$ in $`V`$. Then for $`ϵ>0`$ sufficiently small, $`O_{ϵv}=ϵO_v`$ is contained in $`U`$. This implies that $`\rho `$ is not expansive. On the other hand if $`O_v`$ is unbounded for all non-zero $`vV`$, then it is easy to see that any bounded neighborhood of $`0`$ is an expansive neighborhood. This proves a). Part b) is an immediate consequence of a). $`\mathrm{}`$ ###### Proposition 2.3 Let $`\mathrm{\Gamma }`$ be a semigroup, $`V`$ be a finite-dimensional vector space over $``$ and $`\rho :\mathrm{\Gamma }\mathrm{End}(V)`$ be a reducible endomorphism action with generalized weights $`\lambda _1,\mathrm{},\lambda _k`$. Then $`\rho `$ is expansive if and only if for each $`i`$, the image of $`\lambda _i`$ is an unbounded subset of $``$. #### Proof For $`i=1,\mathrm{},k`$ let $`V_i`$ be the generalized weight space corresponding to $`\lambda _i`$. Since each $`V_i`$ is $`\rho `$-invariant, from the previous proposition it follows that $`\rho `$ is expansive if and only if $`\rho |_{V_i}`$ is expansive for all $`i`$. Therefore without loss of generality we may assume that $`k=1`$. In that case there exists a character $`\lambda `$ of $`\mathrm{\Gamma }`$ such that after suitable identifications we have : a) $`V=^n`$ for some $`n>0`$ and b) For all $`\gamma \mathrm{\Gamma }`$, $`\rho (\gamma )`$ can be written as, $$\rho (\gamma )=\left(\begin{array}{ccccc}\lambda (\gamma )& & & & \\ & \mathrm{}& & & \\ & & \lambda (\gamma )& & \\ & 0& & \mathrm{}& \\ & & & & \lambda (\gamma )\end{array}\right).$$ Suppose there exists a non-zero vector $`v=(v_1\mathrm{},v_n)^t`$ in $`^n`$ such that the $`\rho `$-orbit of $`v`$ is bounded. Let $`jn`$ be the largest positive integer such that $`v_j0`$. Then it is easy to see that $$\{\rho (\gamma )(v)_j|\gamma \mathrm{\Gamma }\}=\{\lambda (\gamma )v_j|\gamma \mathrm{\Gamma }\}.$$ Since the $`\rho `$-orbit of $`v`$ is bounded, this implies that the image of $`\lambda `$ is bounded subset of $``$. Conversely, it is easy to see that if $`v^n`$ denotes the vector $`(1,0,\mathrm{},0)^t`$ then $$\{\rho (\gamma )(v)|\gamma \mathrm{\Gamma }\}=\{(\lambda (\gamma ),0,\mathrm{},0)^t|\gamma \mathrm{\Gamma }\}.$$ Therefore if the image of $`\lambda `$ is bounded then there exists a non-zero vector in $`V`$ with bounded $`\rho `$-orbit. $`\mathrm{}`$ In the special case when $`\mathrm{\Gamma }`$ is isomorphic to $`^+`$ or $``$ we obtain the following consequence of the above proposition. This was first proved in . ###### Proposition 2.4 Let $`V`$ be a finite-dimensional vector space over $``$ or $``$ and let $`T`$ be an element of End$`(V)`$. Then the cyclic semigroup generated by $`T`$ acts expansively on $`V`$ if and only if the spectrum of $`T`$ does not intersect the closed unit disk. If $`T`$ is invertible then the cyclic subgroup generated by $`T`$ acts expansively on $`V`$ if and only if the spectrum of $`T`$ does not intersect $`𝕊^1`$. #### Proof Replacing $`V`$ by $`V`$ if necessary, we may assume that $`V`$ is a complex vector space. Let $`\mathrm{\Gamma }=\{T^n|n0\}`$ denote the cyclic semigroup generated by $`T`$ and let $`\rho `$ denote the natural action of $`\mathrm{\Gamma }`$ on $`V`$. Since $`\mathrm{\Gamma }`$ is abelian, $`\rho `$ is reducible. Furthermore if $`\lambda _1,\mathrm{},\lambda _k`$ denote the generalized weights of $`\rho `$ then $`\lambda _1(T),\mathrm{},\lambda _k(T)`$ are the eigenvalues of $`T`$. Applying the previous proposition we conclude that $`\rho `$ is expansive if and only if the spectrum of $`T`$ does not intersect the closed unit disk. The second assertion follows from a similar argument. $`\mathrm{}`$ ## 3 Expansive actions on Lie groups In this section we prove Theorem A and it’s consequences. Throughout the section $`G`$ will denote a connected Lie group and $`L(G)`$ will denote the Lie algebra of $`G`$. The standard exponential map from $`L(G)`$ to $`G`$ will be denoted by exp. We begin with the following lemma. ###### Lemma 3.1 Let $`V`$ be a finite-dimensional vector space over $``$ and let $`\rho `$ be an expansive endomorphism action of a semigroup $`\mathrm{\Gamma }`$ on $`V`$. Then there exists a finitely generated subsemigroup $`\mathrm{\Gamma }_0\mathrm{\Gamma }`$ which acts expansively on $`V`$. #### Proof Suppose this is not the case. We fix a norm $`||.||`$ on $`V`$. Let $`SV`$ denote the set defined by $`S=\{vV|v=1\}`$. For every finitely generated subsemigroup $`\mathrm{\Lambda }\mathrm{\Gamma }`$ we define a closed subset $`S(\mathrm{\Lambda })S`$ by $$S(\mathrm{\Lambda })=\{vS|\rho (\gamma )(v)1\gamma \mathrm{\Lambda }\}.$$ We claim that for any finitely generated subsemigroup $`\mathrm{\Lambda }\mathrm{\Gamma }`$, $`S(\mathrm{\Lambda })`$ is non-empty. To see this we define a $`\mathrm{\Lambda }`$-invariant subspace $`V(\mathrm{\Lambda })V`$ by $$V(\mathrm{\Lambda })=\{vV|\mathrm{\Lambda }\mathrm{orbit}\mathrm{of}v\mathrm{is}\mathrm{bounded}\}.$$ By Proposition 2.2, $`V(\mathrm{\Lambda })`$ is a non-trivial subspace of $`V`$. Let $`\mathrm{\Lambda }_b\mathrm{End}(V(\mathrm{\Lambda }))`$ be the image of $`\mathrm{\Lambda }`$ under the map $`\gamma \rho (\gamma )|_{V(\mathrm{\Lambda })}`$. Then $`\overline{\mathrm{\Lambda }_b}`$ is a compact subsemigroup of $`\mathrm{End}(V(\mathrm{\Lambda }))`$. We choose a non-zero $`wV(\mathrm{\Lambda })`$ and define a continuous function $`h:\overline{\mathrm{\Lambda }_b}`$ by $`h(\alpha )=\alpha (w)`$ for all $`\alpha `$ in $`\overline{\mathrm{\Lambda }_b}`$. Since $`\overline{\mathrm{\Lambda }_b}`$ is compact, there exists a $`\alpha _0\overline{\mathrm{\Lambda }_b}`$ such that $`h(\alpha _0)h(\alpha )`$ for all $`\alpha `$ in $`\overline{\mathrm{\Lambda }_b}`$. Putting $`v=\alpha _0(w)`$ we see that for any $`\alpha `$ in $`\overline{\mathrm{\Lambda }_b}`$, $$\alpha (v)=h(\alpha \alpha _0)h(\alpha _0)=v.$$ Since $`\overline{\mathrm{\Lambda }_b}`$ contains Id, $`\alpha _0(w)w>0`$. Now it is easy to see that the unit vector $`v/vS(\mathrm{\Lambda })`$, which proves the claim. If $`\mathrm{\Lambda }_1,\mathrm{\Lambda }_2,\mathrm{},\mathrm{\Lambda }_k`$ are finitely generated subsemigroups of $`\mathrm{\Gamma }`$ and $`\mathrm{\Lambda }`$ is a finitely generated subsemigroup of $`\mathrm{\Gamma }`$ containing $`\mathrm{\Lambda }_1,\mathrm{\Lambda }_2,\mathrm{},\mathrm{\Lambda }_k`$, then from the above claim it follows that $$\mathrm{}S(\mathrm{\Lambda })S(\mathrm{\Lambda }_1)\mathrm{}S(\mathrm{\Lambda }_k).$$ This shows that the collection $`\{S(\mathrm{\Lambda })\}`$ has the finite-intersection property. Now from the compactness of $`S`$ it follows that $$\underset{\mathrm{\Lambda }}{}S(\mathrm{\Lambda })\mathrm{}.$$ Clearly for any vector $`w`$ which lies in the intersection, the $`\mathrm{\Gamma }`$-orbit of $`w`$ is bounded, which contradicts our hypothesis. $`\mathrm{}`$ Proof of Theorem A. Suppose $`(G,\rho )`$ is expansive. Let $`U`$ be an expansive neighborhood of the identity in $`G`$. We choose a neighborhood $`V`$ of $`0`$ in $`L(G)`$ such that $`\mathrm{exp}|_V`$ is a homeomorphism and exp($`V`$) is contained in $`U`$. Let $`v`$ be any non-zero element of $`L(G)`$ such that the $`\rho _e`$-orbit of $`v`$ is contained in $`V`$. Since exp is a $`\mathrm{\Gamma }`$-equivariant map from $`(L(G),\rho _e)`$ to $`(G,\rho )`$, it follows that the $`\rho `$-orbit of exp($`v`$) is contained in $`U`$. Since $`U`$ is an expansive neighborhood of the identity and $`\mathrm{exp}|_V`$ is a homeomorphism, this implies that $`v=0`$. Hence $`\rho _e`$ is expansive. Now applying Proposition 2.2 we see that for every non-zero $`v`$ in $`L(G)`$, the $`\rho _e`$-orbit of $`v`$ is unbounded. Now we will prove the converse. Applying Lemma 3.1 and Proposition 2.2 we see that it is enough to consider the case when $`\mathrm{\Gamma }`$ is finitely generated. Let $`A`$ be a finite set which generates $`\mathrm{\Gamma }`$. We choose a neighborhood $`V`$ of $`0`$ in $`L(G)`$ such that $`\mathrm{exp}|_V`$ is a homeomorphism and $`\overline{V}`$ is compact. We choose another neighborhood $`U`$ of $`0`$ such that $$UV\mathrm{and}\rho _e(\gamma )(U)V\gamma A.$$ We claim that exp($`U`$) is an expansive neighborhood of $`(G,\rho )`$. Suppose this is not the case. Let $`ge`$ be any element of $`G`$ such that the $`\rho `$-orbit of $`g`$ is contained in $`\mathrm{exp}(U)`$. We choose $`v`$ in $`U`$ such that $`\mathrm{exp}(v)=g`$. Let $`O_v`$ be the $`\rho _e`$-orbit of $`v`$. Since $`v`$ is non-zero, $`O_v`$ is unbounded. Since $`A`$ generates $`\mathrm{\Gamma }`$ and $`vU`$, it follows that there exists $`w`$ in $`O_v`$ and $`\gamma _0A`$ such that $$wU\text{ and }\rho _e(\gamma _0)(w)VU.$$ Since exp maps the $`\rho _e`$-orbit of $`v`$ onto the $`\rho `$-orbit of exp($`v`$), this implies that the $`\rho `$-orbit of $`\mathrm{exp}(v)`$ intersects $`\mathrm{exp}(VU)`$. Since $`\mathrm{exp}(VU)`$ and $`\mathrm{exp}(U)`$ are disjoint, this gives a contradiction. $`\mathrm{}`$ We note the following consequence of Theorem A and Lemma 3.1. ###### Corollary 3.2 Let $`\mathrm{\Gamma }`$ be a discrete semigroup, $`G`$ be a connected Lie group and $`\rho `$ be an expansive endomorphism action of $`\mathrm{\Gamma }`$ on $`G`$. Then there exists a finitely generated subsemigroup $`\mathrm{\Gamma }_0\mathrm{\Gamma }`$ such that the action of $`\rho `$, restricted to $`\mathrm{\Gamma }_0`$, is expansive. The following corollary provides a rich source of expansive endomorphism actions of non-abelian semigroups on tori. ###### Corollary 3.3 Let $`\mathrm{\Gamma }`$ be an infinite subsemigroup of $`M(n,)`$ which acts irreducibly on $`^n`$. Then the induced $`\mathrm{\Gamma }`$-action on $`𝕋^n^n/^n`$ is expansive. #### Proof Let $`\rho `$ denote the natural action of $`\mathrm{\Gamma }`$ on $`^n`$ and $`W`$ denote the subspace of $`^n`$ consisting of all points with bounded $`\rho `$-orbit. Since $`\mathrm{\Gamma }`$ is an infinite subsemigroup of $`M(n,)`$, it is a noncompact subset of of $`M(n,)`$. Hence $`W`$ is a proper subspace of $`^n`$. From irreducibility of $`\rho `$ we conclude that $`W=\{0\}`$. Now applying Theorem A we see that the action of $`\mathrm{\Gamma }`$ on $`𝕋^n`$ is expansive. $`\mathrm{}`$ ###### Corollary 3.4 Let $`G`$ be a connected Lie group and $`\rho `$ be an expansive automorphism action of a virtually nilpotent group $`\mathrm{\Gamma }`$ on $`G`$. Then $`\rho (\mathrm{\Gamma })`$ contains an expansive automorphism of $`G`$. #### Proof Since $`\mathrm{\Gamma }`$ acts expansively on $`G`$, so does every finite-index subgroup of $`\mathrm{\Gamma }`$. Therefore without loss of generality we may assume that $`\mathrm{\Gamma }`$ is nilpotent. First we will consider the case when $`G`$ is isomorphic to a finite-dimensional vector $`V`$ space over $``$. Let $`HGL(V)`$ be the Zariski-closure of $`\rho (\mathrm{\Gamma })`$ and $`H_0`$ be the connected component of $`H`$ which contains the identity. Since $`H`$ has finitely many components, it follows that $$\mathrm{\Gamma }_0=\{\gamma \mathrm{\Gamma }|\rho (\gamma )H_0\}$$ is a finite-index subgroup of $`\mathrm{\Gamma }`$. Let $`\rho _0`$ denote the $`\mathrm{\Gamma }_0`$-action on $`V`$, induced by $`\rho `$. Since $`\rho (\mathrm{\Gamma }_0)`$ is contained in a connected nilpotent subgroup of $`GL(V)`$, $`\rho _0`$ is reducible. Since $`\rho `$ is expansive, so is $`\rho _0`$. Let $`\lambda _1,\mathrm{},\lambda _k`$ be the generalized weights of $`\rho _0`$. For each $`\gamma `$ in $`\mathrm{\Gamma }_0`$ we define $`A_\gamma \{1,\mathrm{},k\}`$ by $$A_\gamma =\{j||\lambda _i(\gamma )|1\}.$$ For any $`\gamma \mathrm{\Gamma }_0`$ let $`n(\gamma )`$ denote the cardinality of $`A_\gamma `$. We choose $`\gamma _0\mathrm{\Gamma }_0`$ such that the $`n(\gamma _0)n(\gamma )`$ for all $`\gamma `$ in $`\mathrm{\Gamma }_0`$. We claim that $`n(\gamma _0)=k`$ i.e. $`|\lambda _i(\gamma _0)|1`$ for all $`j=1,\mathrm{},k`$. Suppose this is not the case. We choose $`i`$ such that $`|\lambda _i(\gamma _0)|=1`$. By Proposition 2.3 there exists $`\gamma _1`$ such that $`|\lambda _i(\gamma _1)|1`$. It is easy to see that for sufficiently large $`m>0`$, $$|\lambda _j(\gamma _0^m\gamma _1)|1jA_{\gamma _0}\{i\}.$$ Since $`n(\gamma _0)n(\gamma )`$ for all $`\gamma `$ in $`\mathrm{\Gamma }_0`$, this gives a contradiction. Since the numbers $`\lambda _1(\gamma _0),\mathrm{},\lambda _k(\gamma _0)`$ are the eigenvalues of $`\rho (\gamma _0)`$, from the above claim and Proposition 2.4 we see that $`\rho (\gamma _0)`$ is an expansive automorphism of $`V`$. Now we will consider the general case. Let $`\sigma `$ be the $`\mathrm{\Gamma }`$-action on the complexified Lie algebra $`L(G)`$, induced by $`\rho _e`$. Applying Theorem A and Proposition 2.2 we see that $`\sigma `$ is expansive. By the previous argument there exists a $`\gamma _0`$ in $`\mathrm{\Gamma }`$ such that $`\sigma (\gamma _0)`$ is an expansive automorphism of $`L(G)`$. Applying Theorem A and Proposition 2.2 to the cyclic group generated by $`\gamma _0`$, we conclude that $`\rho (\gamma _0)`$ is an expansive automorphism of $`G`$. $`\mathrm{}`$ It is known that if a connected locally compact topological group admits an expansive automorphism then it is nilpotent (see , ). The following corollary generalises this result for actions of virtually nilpotent groups on connected Lie groups. ###### Corollary 3.5 Let $`G`$ be a connected Lie group which admits an expansive automorphism action of a virtually nilpotent group. Then $`G`$ is nilpotent. #### Proof Let $`\rho `$ be an automorphism action of a virtually nilpotent group $`\mathrm{\Gamma }`$ on $`G`$. Then by the previous corollary there exists a $`\gamma \mathrm{\Gamma }`$ such that $`\rho (\gamma )`$ is an expansive automorphism of $`G`$. Applying Theorem A and Proposition 2.4 we see that the spectrum of $`d\gamma \mathrm{Aut}(L(G))`$ does not intersect the unit circle. It is known that if a finite-dimensional Lie algebra over $``$ admits a hyperbolic automorphism then it is nilpotent. Hence $`G`$ is nilpotent. $`\mathrm{}`$ We conclude this section with an example showing that Corollary 3.4 does not hold for actions of arbitrary discrete groups. #### Example 3.6 Fix $`n3`$ and define a subgroup $`\mathrm{\Gamma }`$ of $`GL(n,)`$ by $$\mathrm{\Gamma }=\{\left(\begin{array}{cc}A& b\\ 0& 1\end{array}\right)AGL(n1,),b^{n1}\}$$ It is easy to see that for any $`x=(x_1,\mathrm{},x_n)`$ in $`^n`$, the $`\mathrm{\Gamma }`$-orbit of $`x`$ is given by the set $$\{Ay+x_nb|AGL(n1,),b^{n1}\};y=(x_1,\mathrm{},x_{n1}).$$ Hence for every non-zero $`x`$ in $`^n`$, the $`\mathrm{\Gamma }`$-orbit of $`x`$ is unbounded. Since the spectrum of $`\gamma `$ contains 1 for every $`\gamma `$ in $`\mathrm{\Gamma }`$, applying Proposition 2.4 we see that $`\mathrm{\Gamma }`$ does not contain any expansive automorphism of $`^n`$. ## 4 Expansive actions on solenoids In this section we consider endomorphism actions on solenoids (compact connected finite-dimensional abelian groups). We freely use various results from duality theory of locally compact abelian groups; the reader is referred to for details. For any locally compact abelian group $`G`$, we denote by $`\widehat{G}`$ the dual group of $`G`$. Recall that for a compact connected abelian group $`G`$, we denote by $`L(G)`$ the vector space consisting of all homomorphisms from $`\widehat{G}`$ to $``$, under pointwise addition and scalar multiplication. It is known that if $`G`$ is a solenoid, then $`\widehat{G}`$ is a torsion free discrete abelian group of finite rank. Hence for any solenoid $`G`$, $`L(G)`$ is finite-dimensional. If $`\rho `$ is an endomorphism action of a semigroup $`\mathrm{\Gamma }`$ on a compact connected abelian group $`G`$ then by $`\widehat{\rho }`$ we denote the induced $`\mathrm{\Gamma }`$-action on $`\widehat{G}`$ and by $`\rho _e`$ we denote the induced $`\mathrm{\Gamma }`$-action on $`L(G)`$ defined by $$\rho _e(\gamma )(p)(\chi )=p(\chi \rho (\gamma )).$$ #### Example 4.1 Let $`\mathrm{\Gamma }`$ be a semigroup, $`\widehat{\rho }:\mathrm{\Gamma }M(n,)`$ be a semigroup homomorphism for some $`n1`$ and $`H`$ be a $`\mathrm{\Gamma }`$-invariant subgroup of $`^n`$. Then the action of $`\mathrm{\Gamma }`$ on $`H`$ induces an endomorphism action $`\rho `$ of $`\mathrm{\Gamma }`$ on $`G=\widehat{H}`$. If $`H`$ contains $`^n`$ then $`L(G)`$ can be identified with $`^n`$ and $`\rho _e`$ can be identified with the adjoint of the $`\mathrm{\Gamma }`$-action on $`^n`$ induced by $`\widehat{\rho }`$. From duality theory of compact abelian groups it follows that any endomorphism action of a discrete semigroup on a solenoid can be identified with an action of this form. This section is organized as follows. Throughout this section $`G`$ will denote a solenoid and $`\rho `$ will denote an endomorphism action of a discrete semigroup $`\mathrm{\Gamma }`$ on $`G`$. In 4.1 we prove that the conditions a) and b), as stated in Theorem B are necessary for expansiveness of $`\rho `$. In 4.2 we show that expansiveness of $`\rho `$ and existence of non-trivial bounded $`\rho _e`$-orbit in $`L(G)`$ can be characterized in terms of suitably chosen metrics on $`G`$ and $`L(G)`$ respectively. In 4.3 we prove Theorem B when $`\mathrm{\Gamma }`$ is finitely generated. We complete the proof of Theorem B in 4.4. ### 4.1 Necessary conditions for expansiveness For any compact connected abelian group $`G`$, we define a map $`E`$ from $`L(G)`$ to $`G`$ by the condition $$(\varphi E)(p)=e^{2\pi ip(\varphi )}pL(G),\varphi \widehat{G}.$$ Since for a fixed $`p`$ in $`L(G)`$, the map $`\varphi e^{2\pi ip(\varphi )}`$ is a continuous homomorphism from $`\widehat{G}`$ to $`𝕊^1`$, by the duality theorem the map $`E`$ is well defined and unique. From the uniqueness it follows that $`E`$ is a homomorphism from $`L(G)`$ (considered as an abelian group under addition) to $`G`$. It is easy to check that the kernel of $`E`$ can be identified with the set of all homomorphisms from $`\widehat{G}`$ to $``$, which is a discrete subgroup of $`L(G)`$. #### Remark 4.2 Using the duality theorem we can realize $`L(G)`$ with the set of all one-parameter subgroups of $`G`$ and $`E`$ with the map $`\alpha \alpha (1)`$. In particular when $`G`$ is a torus, $`L(G)`$ can be identified with $`^n`$, the Lie algebra of $`G`$, and $`E`$ can be identified with the standard exponential map. However in general the map $`E`$ is not surjective. In fact from a result of Dixmier (see ) it follows that if $`G`$ is metrizable then $`E`$ is surjective if and only if $`\widehat{G}`$ is a free abelian group. #### Definition Let $`\mathrm{\Gamma }`$ be a semigroup and $`\rho `$ be an endomorphism action of $`\mathrm{\Gamma }`$ on a compact abelian group $`G`$. Then a set $`A\widehat{G}`$ is said to be a $`\rho `$-basis if $`A`$ generates $`\widehat{G}`$ as an abelian group and $`A=_\gamma \widehat{\rho }(\gamma )(F)`$ for some finite set $`F\widehat{G}`$. Clearly $`\widehat{G}`$ admits a $`\rho `$-basis if and only if it is finitely generated as a $`(\mathrm{\Gamma })`$-module. The next two propositions show that for expansive endomorphism actions the conditions a) and b), as stated in Theorem B are satisfied. The results are known, we include the proofs for the sake of completeness. ###### Proposition 4.3 Let $`\mathrm{\Gamma }`$ be a semigroup and $`\rho `$ be an expansive endomorphism action of $`\mathrm{\Gamma }`$ on a compact abelian group $`G`$. Then $`\widehat{G}`$ has a $`\rho `$-basis. #### Proof Let $`U`$ be an expansive neighborhood of of $`e`$ in $`G`$. We choose a finite set $`F\widehat{G}`$ and $`ϵ>0`$ such that $$\underset{\chi F}{}\{gG||\chi (g)1|<ϵ\}U.$$ We define $`A\widehat{G}`$ by $`A=_\gamma \widehat{\rho }(\gamma )(F)`$. Let $`\widehat{H}\widehat{G}`$ be the subgroup generated by $`A`$ and let $`G^{^{}}G`$ be the subgroup consisting of all $`g`$ in $`G`$ such that $`\chi (g)=e`$ for all $`\chi `$ in $`\widehat{H}`$. Then for every $`gG^{^{}}`$, the $`\rho `$-orbit of $`g`$ is contained in $`U`$. Since $`U`$ is an expansive neighborhood of of $`e`$, we conclude that $`G^{^{}}=\{e\}`$. Now from duality theory of compact abelian groups it follows that $`\widehat{H}=\widehat{G}`$ (see , Theorem 53). Hence $`A`$ is a $`\rho `$-basis of $`\widehat{G}`$. $`\mathrm{}`$ ###### Proposition 4.4 (Also see , proof of Corollary 1) Let $`\mathrm{\Gamma }`$ be a semigroup and $`\rho `$ be an expansive endomorphism action of $`\mathrm{\Gamma }`$ on a solenoid $`G`$. Then for every non-zero point $`pL(G)`$, the orbit of $`p`$ under $`\rho _e`$ is unbounded. #### Proof Let $`U`$ be an expansive neighborhood of $`e`$ in $`G`$. Since $`E`$ is a $`\mathrm{\Gamma }`$-equivariant map from $`(L(G),\rho _e)`$ to $`(G,\rho )`$, it follows that $$\underset{\gamma \mathrm{\Gamma }}{}\rho _e(\gamma )^1(E^1(U))E^1(\underset{\gamma \mathrm{\Gamma }}{}\rho (\gamma )^1(U))=\mathrm{Ker}(E).$$ Since the kernel of $`E`$ is discrete, there exists an open set $`VE^1(U)`$ such that $`V\mathrm{ker}(E)=\{0\}`$. From the above identity it is easy to see that $`V`$ is an expansive neighborhood of $`0`$ for the action $`\rho _e`$. Now the given assertion follows from Proposition 2.2. $`\mathrm{}`$ ### 4.2 Metrics induced by endomorphism actions Let $`\delta `$ be the function from $`𝕊^1`$ to $``$ defined by $`\delta (z)=\text{inf}\{|t||t,e^{2\pi it}=z\}.`$ For any solenoid $`G`$ and $`A\widehat{G}`$ we define functions $`d_A:G\times G[0,1]`$ and $`d_A:L(G)\times L(G)[0,\mathrm{}]`$ by $$\begin{array}{ccc}d_A(g,h)\hfill & =\hfill & sup_{\chi A}\delta \chi (g^1h),\hfill \\ d_A^{}(p,q)\hfill & =\hfill & sup_{\chi A}|p(\chi )q(\chi )|.\hfill \end{array}$$ It is easy to see that if $`A`$ generates $`\widehat{G}`$ as an abelian group then $`d_A`$ is a metric on $`G`$ and $`d_A^{}`$ is a metric on the subspace $`L(G)_AL(G)`$ consisting of all $`p`$ in $`L(G)`$ with $`d_A^{}(0,p)<\mathrm{}`$. For $`A\widehat{G}`$ and $`r>0`$, we define open sets $`B_A(r)G`$ and $`B_A^{}(r)L(G)`$ by $$\begin{array}{ccc}B_A(r)\hfill & =\hfill & \{gG|d_A(e,g)<r\},\hfill \\ B_A^{}(r)\hfill & =\hfill & \{pL(G)|d_A^{}(0,p)<r\}.\hfill \end{array}$$ Suppose $`\rho `$ is an endomorphism action of a semigroup $`\mathrm{\Gamma }`$ on a solenoid $`G`$ and $`A`$ is a $`\rho `$-basis of $`\widehat{G}`$. Our next proposition shows that expansiveness of $`\rho `$ and existence of non-trivial bounded $`\rho `$-orbits in $`L(G)`$ can be characterized in terms of the metrics $`d_A`$ and $`d_A^{}`$ respectively. ###### Proposition 4.5 Let $`\mathrm{\Gamma }`$ be a semigroup and $`\rho `$ be an endomorphism action of $`\mathrm{\Gamma }`$ on a solenoid $`G`$. Let $`A`$ be a $`\rho `$-basis of $`\widehat{G}`$. Then we have the following. 1. $`\rho `$ is expansive if and only if $`B_A(ϵ)=\{e\}`$ for some $`ϵ>0`$. 2. $`\rho _e`$ has a nontrivial bounded orbit if and only if $`B_A^{}(C)\{0\}`$ for some $`C>0`$. #### Proof 1) Suppose $`B_A(ϵ)=\{e\}`$ for some $`ϵ>0`$. We choose a finite set $`FA`$ such that $`A=_\gamma \widehat{\rho }(\gamma )(F)`$. Note that for any $`g`$ in $`G`$, the $`\rho `$-orbit of $`g`$ is contained in $`B_F(ϵ)`$ only if $`gB_A(ϵ)`$. Hence $`B_F(ϵ)G`$ is an expansive neighborhood of $`e`$ for the action $`\rho `$. On the other hand suppose $`UG`$ is an expansive neighborhood of $`e`$. We choose a finite set $`FA`$ such that $`A=_\gamma \widehat{\rho }(\gamma )(F)`$. Since $`A`$ generates $`\widehat{G}`$ as an abelian group, it follows that $`B_F(ϵ)U`$ for some $`ϵ>0`$. Now it is easy to see that $$B_A(ϵ)=\underset{\gamma }{}\rho (\gamma )^1(B_F(ϵ))\underset{\gamma }{}\rho (\gamma )^1(U)=\{e\}.$$ 2) Suppose $`p`$ is a non-zero element of $`L(G)`$ such that the $`\rho _e`$-orbit of $`p`$ is bounded. We fix an element $`\varphi `$ in $`\widehat{G}`$ and observe that $$\{q(\varphi )|q\rho _e(\mathrm{\Gamma })(p)\}=\{p(\psi )|\psi \widehat{\rho }(\mathrm{\Gamma })(\varphi )\}.$$ Since $`qq(\varphi )`$ is a continuous map from $`L(G)`$ to $``$ and the $`\rho _e`$-orbit of $`p`$ is bounded, the right hand side is a bounded subset of $``$. Hence $`p`$, restricted to the $`\widehat{\rho }`$-orbit of $`\varphi `$ is bounded. Since $`\varphi `$ is arbitrary and $`A`$ is a union of finitely many $`\widehat{\rho }`$-orbits, we conclude that $`p|_A`$ is bounded i.e. $`pB_A^{}(C)`$ for some $`C>0`$. Conversely, suppose $`p`$ is an element of $`B_A^{}(C)`$ for some $`C>0`$. We fix an element $`\varphi `$ in $`\widehat{G}`$. Since $`A`$ generates $`\widehat{G}`$ as an abelian group, there exists a positive integer $`l`$ and $`a_1,\mathrm{},a_l`$ in $`A`$ such that $`\varphi =a_1+\mathrm{}+a_l`$. Since $`A`$ is invariant under $`\widehat{\rho }`$, every element in the $`\widehat{\rho }`$-orbit of $`\varphi `$ can be written as sum of $`l`$ elements of $`A`$. Hence $`|p(\psi )|<lC`$ for all $`\psi `$ in the $`\widehat{\rho }`$-orbit of $`\varphi `$ i.e. $`|q(\varphi )|<lC`$ for all $`q`$ in the $`\rho _e`$-orbit of $`p`$. Since $`\varphi `$ is arbitrary, this implies that the $`\rho _e`$-orbit of $`p`$ is bounded. $`\mathrm{}`$ ### 4.3 Actions of finitely generated semigroups As remarked earlier, for an arbitrary solenoid $`G`$, the map $`E:L(G)G`$ need not be surjective. However we will prove the following result which will be a crucial step in the proof of Theorem B. ###### Theorem 4.6 Let $`\mathrm{\Gamma }`$ be a finitely generated semigroup and $`\rho `$ be an endomorphism action of $`\mathrm{\Gamma }`$ on a solenoid $`G`$. Let $`A`$ be a $`\rho `$-basis of $`\widehat{G}`$. Then for any $`C>0`$ there exists an $`ϵ>0`$ such that $`B_A(ϵ)E(B_A^{}(C))`$. Note that this theorem, together with Propositions 4.3, 4.4 and 4.5 implies Theorem B when $`\mathrm{\Gamma }`$ is finitely generated. #### Remark 4.7 Earlier we have observed that $`E`$ is a continuous homomorphism and the kernel of $`E`$ is discrete. Hence the above theorem implies that the map $`E`$, restricted to $`L(G)_A`$, is a local homeomorphism if $`L(G)_A`$ and $`G`$ are equipped with the topologies induced by $`d_A^{}`$ and $`d_A`$ respectively. Before beginning the proof of Theorem 4.6 we introduce a few notations. Let $`H`$ be a discrete abelian group and $`A`$ is a finite subset of $`H`$. Then by $`𝐐(A)`$ we denote the set of all $`hH`$ which satisfies a linear equation of the form $$n_0h=\underset{j=1}{\overset{r}{}}n_ja_j,$$ where $`a_1,\mathrm{},a_r`$ are elements of $`A`$ and $`n_0,n_1,\mathrm{},n_r`$ are integers with $`n_00`$. Also for all $`k>0`$, by $`𝐐_k(A)`$ we denote the set of all $`hH`$ which satisfies a linear equation as above with $`|n_0|+|n_1|+\mathrm{}+|n_r|k`$. A set $`AH`$ is said to be a $`k`$-regular if there exists an increasing sequence of finite sets $`A_1A_2\mathrm{}A`$ satisfying $$\underset{i=1}{\overset{\mathrm{}}{}}A_i=A,A_n𝐐_k(A_{n1})n2.$$ Now we show that if $`\rho `$ is an endomorphism action of a finitely generated semigroup on a solenoid $`G`$, then any $`\rho `$-basis is $`k`$-regular for some $`k`$. ###### Proposition 4.8 Let $`\mathrm{\Gamma }`$ be a finitely generated semigroup and $`\rho `$ be an endomorphism action of $`\mathrm{\Gamma }`$ on a solenoid $`G`$. Then any $`\rho `$-basis of $`\widehat{G}`$ is $`k`$-regular for some $`k>0`$. #### Proof Let $`A`$ be a $`\rho `$-basis of $`\widehat{G}`$. Since $`A`$ generates $`\widehat{G}`$ as an abelian group and $`\widehat{G}`$ has finite rank, there exists a finite set $`F_1A`$ such that $`𝐐(F_1)=\widehat{G}`$. Also, $`A=_\gamma \widehat{\rho }(\gamma )(F_2)`$ for some finite set $`F_2A`$. We define $`F=F_1F_2`$. Let $`S\mathrm{\Gamma }`$ be finite set which generates $`\mathrm{\Gamma }`$ as a semigroup. Since $`𝐐(F)=\widehat{G}`$, for any $`\gamma `$ in $`S`$ and $`a_0`$ in $`F`$, $`\widehat{\rho }(\gamma )(a_0)`$ satisfies a linear equation of the form $$n_0\widehat{\rho }(\gamma )(a_0)=\underset{j=1}{\overset{r}{}}n_ja_j,$$ where $`a_1,\mathrm{},a_r`$ are elements of $`F`$ and $`n_0,n_1,\mathrm{},n_r`$ are integers with $`n_00`$. Since $`S`$ is finite, it follows that there exists a positive integer $`k`$ such that $`\widehat{\rho }(\gamma )(F)𝐐_k(F)`$ for all $`\gamma `$ in $`S`$. For $`n1`$ we define $`S_n\mathrm{\Gamma }`$ and $`A_nA`$ by $$S_n=\{\gamma _1\mathrm{}\gamma _n|\gamma _1,\mathrm{},\gamma _nS\},A_n=\underset{\gamma S_n}{}\widehat{\rho }(\gamma )(F).$$ Let $`\gamma _n`$ be any element of $`S_n`$. We choose $`\gamma _{n1}S_{n1}`$ and $`\gamma _0S`$ such that $`\gamma _n=\gamma _{n1}\gamma _0`$. Since $`\widehat{\rho }(\gamma _0)(F)𝐐_k(F)`$, we see that $$\widehat{\rho }(\gamma _n)(F)\widehat{\rho }(\gamma _{n1})(𝐐_k(F))𝐐_k(\widehat{\rho }(\gamma _{n1})(F)).$$ This shows that $`A_n𝐐_k(A_{n1})`$ for all $`n2`$. Since $`S`$ generates $`\mathrm{\Gamma }`$, it is easy to see that $`A_i=\widehat{\rho }(\mathrm{\Gamma })(F)=A`$. $`\mathrm{}`$ Now we turn to the proof of Theorem 4.6. We will use the following lemma. ###### Lemma 4.9 Let $`G`$ be a solenoid and $`F`$ be a finite subset of $`\widehat{G}`$. Then for any $`C>0`$ there exists an $`ϵ>0`$ such that for any $`gB_F(ϵ)`$ there exists $`p`$ in $`B_F^{}(C)`$ satisfying $`\varphi (g)=\varphi E(p)`$ for all $`\varphi `$ in $`F`$. #### Proof Let $`\widehat{H}\widehat{G}`$ be the subgroup generated by $`F`$ and let $`H`$ be the dual of $`\widehat{H}`$. Since $``$ is divisible, any homomorphism from $`\widehat{H}`$ to $``$ can be extended to a homomorphism from $`\widehat{G}`$ to $``$. Therefore it is enough to consider the case when $`G=H`$. In that case there exists $`n1`$ and an isomorphism $`\theta :G𝕋^n`$ such that $`\mathrm{exp}\mathrm{d}\theta =\theta E`$. Since $`B_F^{}(C)`$ is an open subset of $`L(G)`$ and $`\mathrm{exp}:L(𝕋^n)𝕋^n`$ is a local homeomorphism, it is easy to see that $`E(B_F^{}(C))`$ contains an open neighborhood of $`e`$ in $`G`$. Since $`F`$ generates $`\widehat{G}`$, there exists an $`ϵ>0`$ such that $`B_F(ϵ)E(B_F^{}(C))`$. This proves the lemma. $`\mathrm{}`$ Proof of Theorem 4.6 Suppose $`A`$ is a $`\rho `$-basis of $`\widehat{G}`$. By Proposition 4.8 $`A`$ is $`k`$-regular for some $`k>0`$. Clearly without loss of generality we may assume that $`C<1/k`$. Let $`A_1A_2\mathrm{}A`$ be an increasing sequence of finite sets such that $$\underset{i}{}A_i=A,A_n𝐐_k(A_{n1})n2.$$ Applying Lemma 4.9 we choose a positive $`ϵ<C`$ such that for all $`hB_{A_1}(ϵ)`$ there exists a $`qB_{A_1}^{}(C)`$ satisfying $`\chi (E(q))=\chi (h)`$ for all $`\chi `$ in $`A_1`$. Let $`g`$ be an element of $`B_A(ϵ)`$. We choose a $`p`$ in $`B_{A_1}^{}(C)`$ such that $`\chi (E(p))=\chi (g)`$ for all $`\chi `$ in $`A_1`$. We claim that $$|p(\chi )|<C,\chi (E(p))=\chi (g)\chi A.$$ By our choice of $`p`$ this is true for all $`\chi `$ in $`A_1`$. Suppose this holds for all $`\chi `$ in $`A_n`$. Let $`a_0`$ be an element of $`A_{n+1}`$. We choose integers $`n_0,n_1,\mathrm{},n_r`$ and $`a_1,\mathrm{},a_rA_n`$ such that $`n_00`$, $`|n_0|+|n_1|+\mathrm{}+|n_r|k`$ and $$n_0a_0=\underset{j=1}{\overset{r}{}}n_ja_j.$$ Since $`gB_A(ϵ)`$, there exists $`\alpha `$ such that $`|\alpha |<ϵ<C`$ and $`e^{2\pi i\alpha }=a_0(g)`$. Now from the above equation it is easy to see that $$e^{2\pi in_0\alpha }=a_0(g)^{n_0}=\underset{1}{\overset{r}{}}a_i(g)^{n_i}.$$ Since $`a_1,\mathrm{},a_rA_n`$, it follows that $$e^{2\pi in_0p(a_0)}=\underset{1}{\overset{r}{}}e^{2\pi in_jp(a_j)}=\underset{1}{\overset{r}{}}a_iE(p)^{n_i}=\underset{1}{\overset{r}{}}a_i(g)^{n_i}.$$ Therefore $`e^{2\pi in_0(\alpha p(a_0))}=0`$ i.e. $`n_0(\alpha p(a_0))`$. On the other hand $$|n_0(\alpha p(a_0))|n_0|\alpha |+|p(n_0a_0)|n_0C+\underset{1}{\overset{r}{}}n_i|p(a_i)|.$$ Now applying our induction hypothesis we see that $$|n_0(\alpha p(a_0))|C\underset{j=1}{\overset{r}{}}|n_j|<Ck<\mathrm{\hspace{0.17em}1}.$$ Hence $`\alpha p(a_0)=0`$, which implies that $`|p(a_0)|=|\alpha |<C`$ and $`a_0(E(p))=e^{2\pi i\alpha }=a_0(g)`$. This proves the claim. Since $`A`$ generates $`\widehat{G}`$ as an abelian group, from the above claim we deduce that for any $`gB_A(ϵ)`$ there exists a $`pB_A^{}(C)`$ such that $`E(p)=g`$. This completes the proof. $`\mathrm{}`$ ### 4.4 Proof of Theorem B Now we prove Theorem B for an endomorphism action $`\rho `$ of an arbitrary semigroup $`\mathrm{\Gamma }`$ on a solenoid $`G`$. If $`(G,\rho )`$ is expansive then from Proposition 4.3 and Proposition 4.4 it follows that a) and b) are satisfied. Conversely, suppose the conditions a) and b) are satisfied. By Lemma 3.1 there exists a finitely generated subsemigroup $`\mathrm{\Gamma }_0`$ of $`\mathrm{\Gamma }`$ such that for all $`p`$ in $`L(G)`$, the $`\mathrm{\Gamma }_0`$-orbit of $`p`$ under the action $`\rho _e`$ is unbounded. We construct an endomorphism action $`(H,\sigma )`$ of $`\mathrm{\Gamma }_0`$ as follows : Since $`\widehat{G}`$ has finite rank, there exists a finite set $`F_1\widehat{G}`$ such that $`𝐐(F_1)=\widehat{G}`$. Also, from a) it follows that there exists a finite set $`F_2\widehat{G}`$ such that the set $`_\gamma \widehat{\rho }(\gamma )(F_2)`$ generates $`\widehat{G}`$ as an abelian group. We define $`F=F_1F_2`$. Let $`\widehat{H}\widehat{G}`$ be the subgroup generated by the set $$A_0=\underset{\gamma \mathrm{\Gamma }_0}{}\widehat{\rho }(\gamma )(F)$$ and let $`\widehat{\sigma }`$ be the $`\mathrm{\Gamma }_0`$-action on $`\widehat{H}`$ defined by $`\widehat{\sigma }(\gamma )=\widehat{\rho }(\gamma )|_{\widehat{H}}`$. Let $`H`$ denote the dual of $`\widehat{H}`$ and let $`\sigma `$ denote the endomorphism action of $`\mathrm{\Gamma }_0`$ on $`H`$ which is the dual of $`\widehat{\sigma }`$. Claim : $`(H,\sigma )`$ is expansive. Since $`\widehat{G}=𝐐(F)𝐐(\widehat{H})`$, it follows that the restriction map $`pp|_{\widehat{H}}`$ is a $`\mathrm{\Gamma }_0`$-equivariant linear isomorphism from $`L(G)`$ onto $`L(H)`$. This implies that for every non-zero $`qL(H)`$, the $`\sigma _e`$-orbit of $`q`$ is unbounded in $`L(H)`$. Also it is easy to see that $`A_0`$ is a $`\sigma `$-basis of $`\widehat{H}`$. Now applying Theorem 4.6 and Proposition 4.5 we see that $`(H,\sigma )`$ is expansive, which proves the claim. Let $`V`$ be an expansive neighborhood of $`e`$ in $`H`$. Let $`i:\widehat{H}\widehat{G}`$ be the inclusion map and let $`\pi :GH`$ be the dual of $`i`$. We claim that $`\pi ^1(V)`$ is an expansive neighborhood for the action $`\rho `$. To see this we choose any $`g`$ in $`G`$ such that the $`\rho `$-orbit of $`g`$ is contained in $`\pi ^1(V)`$. It is easy to see that for any $`\gamma \mathrm{\Gamma }_0`$, $$\sigma (\gamma )\pi (g)=\pi \rho (\gamma )(g).$$ Therefore the $`\sigma `$-orbit of $`\pi (g)`$ is contained in $`V`$. Since $`V`$ is an expansive neighborhood for $`\sigma `$, this implies that $`\pi (g)=e`$. In particular $`\varphi (g)=1`$ for all $`\varphi `$ in $`F`$. For any $`\gamma `$ in $`\mathrm{\Gamma }`$, replacing $`g`$ by $`\rho (\gamma )(g)`$ and applying the same argument we see that $$\widehat{\rho }(\gamma )(\varphi )(g)=\varphi \rho (\gamma )(g)=1\gamma \mathrm{\Gamma },\varphi F.$$ Since $`A=_{\gamma \mathrm{\Gamma }}\widehat{\rho }(\gamma )(F)`$ generates $`\widehat{G}`$, it follows that $`g=e`$. This completes the proof. $`\mathrm{}`$ #### Remark 4.10 Let $`\mathrm{\Gamma }`$ be the group of positive rational numbers under multiplication, $``$ be the group of rational numbers under addition and $`\widehat{\rho }`$ be the action of $`\mathrm{\Gamma }`$ on $``$ defined by $`\widehat{\rho }(\gamma )(x)=\gamma x`$. Let $`\rho `$ be the dual action of $`\mathrm{\Gamma }`$ on $`\widehat{}`$. Then from Theorem B it follows that $`(\widehat{},\rho )`$ is expansive. On the other hand it is easy to see that for any finitely generated subgroup $`\mathrm{\Gamma }_0\mathrm{\Gamma }`$, $``$ is not finitely generated as a $`(\mathrm{\Gamma }_0)`$-module. Hence no finitely generated subgroup $`\mathrm{\Gamma }_0\mathrm{\Gamma }`$ acts expansively on $`\widehat{}`$ under the action $`\rho `$. Since $`\mathrm{\Gamma }`$ is abelian, this shows that analogues of Corollary 3.2 and Corollary 3.4 do not hold when $`G`$ is a solenoid. Address : School of Mathematics, Tata Institute of Fundamental Research, Mumbai 400005, India. E-mail : siddhart@math.tifr.res.in
warning/0003/math0003058.html
ar5iv
text
# Quantum co-adjoint orbits of MD₄-groups. ## 1. Introduction First of all, we recall the notion of K-action (see \[Ki1\]). Let us denote by $`G`$ a connected and simply connected Lie group, its Lie algebra $`𝔤=T_eG`$ as the tangent space at the neutral element e. It is easy to see that to each element $`gG`$ one can associate a map $$A(g):GG$$ by conjugacy, in fixing the neutral element $`eG`$. Therefore, there is a corresponding tangent map $$A(g)_{}:𝔤=T_eG𝔤=T_eG$$ $$X𝔤\frac{d}{dt}g\mathrm{exp}(tX)g^1|_{t=0}𝔤.$$ This defines an action, denoted as usually by $`Ad`$, of group $`G`$ in its Lie algebra $`𝔤`$. We define the co-adjoint action of group $`G`$ in the dual vector space $`𝔤^{}`$ by the formula: $$K(g)F,X:=F,Ad(g^1)X$$ for all $`F𝔤^{},X𝔤,gG`$. It is easy to check that this defines an action of $`G`$ on $`𝔤^{}.`$ As an easy consequence, the dual space $`𝔤^{}`$ is decomposed into a disconnected sum of the K-orbits. In 1980 Do Ngoc Diep introduced the notion of the $`MD`$-groups, $`MD`$-algebras and then Le Anh Vu gave a complete classification of the $`MD_4`$-groups (see \[D\]). In \[DH1\], \[DH2\] and \[H3\] we obtained the exact formulae of deformation quantization and therefore performed the corresponding quantun coadjoint orbits for the $`\overline{MD}`$-groups (i.e. the groups, every K-orbit of which is of dimension, equal 0 or dim $`G`$) and for the real diamond Lie group. In this article, applying the procedure of deformation quantization we shall obtain quantum co-adjoint orbits of all the $`MD_4`$-groups. The article is organized as follows. In section 2 we recall the basic definitions, preliminary results. Each adapted chart that carries the Moyal $``$-product from $`^2`$ onto co-adjoint orbits $`\mathrm{\Omega }_F`$ of the exponential $`MD_4`$-groups, in particular, Hamiltonian functions in canonical coordinates, are introduced in section 3. Section 4 is devoted to the following groups $$G_{4,2,3(\frac{\pi }{2})};G_{4,2,4};G_{4,3,4(\frac{\pi }{2})};G_{4,4,1}.$$ By direct computations and by exponentiating we obtain the corresponding unitary representations of the $`MD_4`$-groups. ## 2. Basic definitions and Preliminary results ###### DEFINITION 2.0.1. (\[D\]) We say that a solvable Lie group $`G`$ belongs to the class $`MD`$ if and only if every its K-orbit has dimension 0 or maximal. A Lie algebra is of class $`MD`$ if and only if its corresponding Lie group is of the same class. We also recall the following results for $`MD_4`$-algebras (i.e. $`dim𝔤=4`$ ): ###### Theorem 2.0.2. Assume $`𝔤`$ is a $`MD_4`$-algebra with generators $`X,Y,Z,T`$, then * If $`𝔤`$ is decomposable then it is of the form $$𝔤=^n\stackrel{~}{𝔤}$$ for n=1,2,3,4 and some indecomposable ideal $`\stackrel{~}{𝔤}.`$ * If $`𝔤`$ is indecomposable then $`𝔤`$ is of class $`MD_4`$ if and only if it is generated by the generators $`X,Y,Z,T`$ with only non-trivial commutation relation which is one of following relations defined in each case: * $`𝔤^1=[𝔤,𝔤]=Z,`$ and + $`[T,X]=Z`$ $`(𝔤_{4,1,1})`$ + $`[T,Z]=Z`$ $`(𝔤_{4,1,2})`$ * $`𝔤^1=[𝔤,𝔤]=Y+Z^2`$, and + $`[T,Y]=\lambda Y,[T,Z]=Z;\lambda ^{}=\backslash (0)(𝔤_{4,2,1(\lambda )})`$ + $`[T,Y]=Y;[T,Z]=Y+Z(𝔤_{4,2,2})`$ + $`ad_T=\left(\begin{array}{ccc}cos\phi & sin\phi & 0\\ sin\phi & cos\phi & 0\\ 0& 0& 0\end{array}\right)`$, $`(𝔤_{4,2,3(\phi )})`$ + $`ad_T=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 0\end{array}\right),`$ $`ad_X=\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)`$ $`(𝔤_{4,2,4})=Lie(Aff())`$ * $`𝔤^1=[𝔤,𝔤]=X+Y+Z`$, and + $`ad_T=\left(\begin{array}{ccc}\lambda _1& 0& 0\\ 0& \lambda _2& 0\\ 0& 0& 1\end{array}\right)`$, $`\lambda _1,\lambda _2^{}`$ $`(𝔤_{4,3,1(\lambda _1,\lambda _2)})`$ + $`ad_T=\left(\begin{array}{ccc}\lambda & 1& 0\\ 0& \lambda & 0\\ 0& 0& 1\end{array}\right)`$,$`\lambda ^{}`$ $`(𝔤_{4,3,2(\lambda )})`$ + $`ad_T=\left(\begin{array}{ccc}1& 1& 0\\ 0& 1& 1\\ 0& 0& 1\end{array}\right)`$ $`(𝔤_{4,3,3})`$ + $`ad_T=\left(\begin{array}{ccc}cos\phi & sin\phi & 0\\ sin\phi & cos\phi & 0\\ 0& 0& \lambda \end{array}\right)`$,$`\lambda ^{},\phi (0,\pi )`$ $`(𝔤_{4,3,4(\lambda )})`$ * $`𝔤^1=[𝔤,𝔤]=X+Y+Z𝔥_3`$ -the 3-dimensional Heisenberg Lie algebra, and + $`ad_T=\left(\begin{array}{ccc}0& 1& 0\\ 1& 0& 0\\ 0& 0& 0\end{array}\right)`$, $`[X,Y]=Z`$ $`(𝔤_{4,4,1}=Lie(_j_3))`$ + $`ad_T=\left(\begin{array}{ccc}1& 0& 0\\ 0& 1& 0\\ 0& 0& 0\end{array}\right)`$, $`[X,Y]=Z`$ $`(𝔤_{4,4,2}=Lie(_3))`$ (In this case the group is called the real diamond Lie group, see\[H3\]) Until now we fix a $`MD_4`$\- algebra $`𝔤`$ with the standard basis X,Y,Z,T. It is isomorphic to $`^4`$ as vector spaces. The coordinates in this standard basis is denoted by (a,b,c,d). We identify its dual vector spase $`𝔤^{}`$ with $`^4`$ with the help of the dual basis $`X^{},Y^{},Z^{},T^{}`$ and with the local coordinates $`(\alpha ,\beta ,\gamma ,\delta )`$. Thus, for all $`U𝔤,U=aX+bY+cZ+dT`$ and for all $`F𝔤^{},F=\alpha X^{}+\beta Y^{}+\gamma Z^{}+\delta T^{}.`$ Finally, $`\mathrm{\Omega }_F`$ is the co-adjoint orbit passing through $`F𝔤^{}.`$ ###### Theorem 2.0.3. (The Picture of Co-adjoint Orbit)\[D\] * Case $`G=G_{4,1,1}`$ + Each point F with the coordinate $`\gamma =0`$ is a 0-dimensional co-adjoint orbit: $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,\beta ,0,\delta )}$$ + The subset $`\gamma 0`$ is decomposed into a family of 2-dimensional co-adjoint orbits: (1) $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\beta ,\gamma 0)}=\{(\alpha +\gamma d,\beta ,\gamma ,\gamma a+\delta )\}=\{(x,\beta ,\gamma ,t)|x,t\},$$ which are planes. * Case $`G=G_{4,1,2}`$ + Each point F with the coordinate $`\gamma =0`$ is a 0-dimensional co-adjoint orbit: $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,\beta ,0,\delta )}$$ + The subset $`\gamma 0`$ is decomposed into a family of 2-dimensional co-adjoint orbits: (2) $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,\beta )}=\{\alpha ,\beta ,\gamma e^d,\gamma c\underset{1}{\overset{\mathrm{}}{}}\frac{d^{n1}}{n!}+\delta \}=\{(\alpha ,\beta ,z,t)|z,t,\gamma z>0\}$$ which are half-planes, parameterized by the coordinates $`\alpha ,\beta `$ * Case $`G=G_{4,2,1(\lambda )},\lambda ^{}`$ + Each point on the plane $`\beta =\gamma =0`$ is a 0-dimensional co-adjoint orbit: $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,0,0,\delta )}$$ + The open set $`\beta ^2+\gamma ^20`$ is decomposed into the union of 2-dimensional cylinders (3) $$\mathrm{\Omega }_F=\{(\alpha ,\beta e^{s\lambda },\gamma e^s,t)|s,t\}$$ * Case $`G=G_{4,2,2}`$ + Each point on the plane $`\beta =\gamma =0`$ is a 0-dimensional co-adjoint orbit: $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,0,0,\delta )}$$ + The open set $`\beta ^2+\gamma ^20`$ is decomposed into the union of 2-dimensional cylinders: (4) $$\mathrm{\Omega }_F=\{(\alpha ,\beta e^s,\beta se^s+\gamma e^s,t)|s,t\}$$ * Case $`G=G_{4,2,3(\phi )}`$ with $`\phi (0,\pi ).`$ We identify $`𝔤_{4,2,3(\phi )}^{}`$ with $`\times \times `$ and $`F=(\alpha ,\beta ,\gamma ,\delta )`$ with $`(\alpha ,\beta +i\gamma ,\delta ).`$ Then, + Each point $`(\alpha ,0,\delta )`$ is a 0-dimensional co-adjoint orbit: $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,0+i0,\delta )}$$ + The open set $`\beta +i\gamma 0`$ is decomposed into the union of 2-dimensional co-adjoint orbits: (5) $$\mathrm{\Omega }_F=\{(\alpha ,(\beta +i\gamma )e^{se^{i\phi }},t)|s,t\},$$ which are also cylinders. * Case $`G=G_{4,2,4}=\stackrel{~}{Aff}()`$ + Each point $`(\alpha ,0,0,\delta )`$ is a 0-dimensional co-adjoint orbit: $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\alpha ,0,0,\delta )}$$ + The open set $`\beta ^2+\gamma ^20`$ is the single 4-dimensional co-adjoint orbit: (6) $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(\beta ^2+\gamma ^2)0}=\{(x,y,z,t)|y^2+z^20\}=\times (^2)^{}\times $$ * Case $`G`$ is one of the groups $`G_{4,3,1(\lambda _1,\lambda _2)},`$ $`G_{4,3,2(\lambda )}`$ or $`G_{4,3,3}`$ + Each point $`F=\delta T^{}`$ on the line $`\alpha =\beta =\gamma =0`$ is a 0-dimensional co-adjoint orbit. + The open set $`\alpha ^2+\beta ^2+\gamma ^20`$ is decomposed into a family of co-adjoint orbits which are cylinders, coressponding to the groups $`G_{4,3,1(\lambda _1,\lambda _2)}`$,$`G_{4,3,2(\lambda )}`$ or $`G_{4,3,3}`$: (7) $$\mathrm{\Omega }_F=\{(\alpha e^{s\lambda _1},\beta e^{s\lambda _2},\gamma e^s,t)|s,t\}.$$ (8) $$\mathrm{\Omega }_F=\{(\alpha e^{s\lambda },\alpha se^{s\lambda }+\beta e^{s\lambda },\gamma e^s,t)|s,t\}.$$ (9) $$\mathrm{\Omega }_F=\{(\alpha e^{s\lambda },\alpha se^s+\beta e^s,\frac{1}{2}\alpha s^2e^s+\beta se^s+\gamma e^s,t)|s,t\}.$$ * Case $`G=G_{4,3,4(\lambda ,\phi )}`$ for $`\lambda ^{}`$, $`\phi (0,\pi ).`$ We identify $`𝔤_{4,3,4(\lambda ,\phi )}^{}`$ with $`\times ^2`$ and $`F=(\alpha ,\beta ,\gamma ,\delta )`$ with $`(\alpha +i\beta ,\gamma ,\delta ).`$ Then, + Each point of the line defined by the condition $`\alpha =\beta =\gamma =0`$ is a 0-dimensional co-adjoint orbit $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(0,0,\delta )}=\{(0+i.0,0,\delta )\}.$$ + The open set $`|\alpha +i\beta |^2+\gamma ^20`$ is decomposed into an union of co-adjoint orbits, which are cylinders (10) $$\mathrm{\Omega }_F=\{((\alpha +i\beta )e^{se^{i\phi }},\gamma e^{s\lambda },t)|s,t\}.$$ * Case $`G=G_{4,4,1}=_j𝔥_3`$ + Each point of the line defined by the conditions $`\alpha =\beta =\gamma =0`$ is a 0-dimensional co-adjoint orbit $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(0,0,0,\delta )}=\{(0,0,0,\delta )\}.$$ + The set $`\alpha ^2+\beta ^20,\gamma =0`$ is the union of 2- dimensional co-adjoint orbits, which are rotation cylinders (11) $$\mathrm{\Omega }_F=\{(\alpha cos\theta \beta sin\theta ,\alpha sin\theta +\beta cos\theta ,0,t)|\theta ,t\}$$ i.e $$\mathrm{\Omega }_F=\{(x,y,0,t)|x^2+y^2=\alpha ^2+\beta ^2;x,y,t\}.$$ + The open set $`\gamma 0`$ is decomposed into a union of 2-dimensional co-adjoint orbits (12) $$\mathrm{\Omega }_F=\{(x,y,\gamma ,t)|x^2+y^22\gamma t=\alpha ^2+\beta ^22\gamma \delta ;x,y,t\},$$ which are rotation paraboloids * Case $`G=G_{4,4,2}=_3`$, the real diamond group (see \[H3\]) + Each point of the line $`\alpha =\beta =\gamma =0`$ is a 0-dimensional co-adjoint orbit $$\mathrm{\Omega }_F=\mathrm{\Omega }_{(0,0,0,\delta )}$$ + The set $`\alpha 0,\beta =\gamma =0`$ is union of 2-dimensional co-adjoint orbits ,which are just half-planes (13) $$\mathrm{\Omega }_F=\{(x,0,0,t)|x,t,\alpha x>0\}$$ + The set $`\alpha =\gamma =0,\beta 0`$ is union of 2-dimensional co-adjoint orbits, which are just half-planes (14) $$\mathrm{\Omega }_F=\{(0,y,0,t)|y,t,\beta y>0\}.$$ + The set $`\alpha \beta 0,\gamma =0`$ is decomposed into a family of 2-dimensional co-adjoint orbits, which are just hyperbolic-cylinders (15) $$\mathrm{\Omega }_F=\{(x,y,0,t)|x,y,t\&\alpha x>0,\beta y>0,xy=\alpha \beta \}.$$ + The open set $`\gamma 0`$ is decomposed into a family of 2-dimensional co-adjoint orbits , which are just hyperbolic- paraboloids (16) $$\mathrm{\Omega }_F=\{(x,y,\gamma ,t)|x,y,t\&xy\alpha \beta =\gamma (t\delta )\}.$$ Thus, we have 15 family of 2-dimensional co-adjoint orbits and a 4-dimensional co-adjoint orbit $`\mathrm{\Omega }_F\times ^{}`$. They are strictly homogeneous symplectic manifolds with a flat action. Deformation of Poisson brackets and associative algebras of $`C^{\mathrm{}}`$-functions on symplectic manifold (classical phase spaces) permit an autonomous quantization theory without need for Hilbert space operators. This theory is based on the of $``$-products introduced by Flato and Lichnerowicz (see for instance \[AC1\]). Let us denote by $`\mathrm{\Lambda }`$ the 2-tensor associated with the Kirillov standard form $`\omega =dpdq`$ in canonical Darboux coordinates. We consider the well-known Moyal $``$-product of two smooth functions $`u,vC^{\mathrm{}}(^2)`$ ( see e.g \[DH1\], \[H3\]), defined by $$uv=u.v+\underset{r1}{}\frac{1}{r!}(\frac{1}{2i})^rP^r(u,v),$$ where $$P^1(u,v)=\{u,v\}$$ $$P^r(u,v):=\mathrm{\Lambda }^{i_1j_1}\mathrm{\Lambda }^{i_2j_2}\mathrm{}\mathrm{\Lambda }^{i_rj_r}_{i_1i_2\mathrm{}i_r}^ru_{j_1j_2\mathrm{}j_r}^rv,$$ with $$_{i_1i_2\mathrm{}i_r}^r:=\frac{^r}{x^{i_1}\mathrm{}x^{i_r}};x:=(p,q)=(p_1,\mathrm{},p_n,q^1,\mathrm{},q^n).$$ It is well-known that this series converges in the Schwartz distribution spaces $`𝒮(^n)`$. Remark that the $`MD_4`$-groups are not all nilpotent or exponential groups. In the most general context, some quantum co-adjoint orbits appeared in \[AC1\]\[AC2\]. However, it is difficult to calculate precisely the $``$-product in concrete cases. In this article we will give explicit formulas for co-adjoint orbits of all $`MD_4`$-groups, even for groups which are neither nilpotent nor exponential. ## 3. Quantum co-adjoint orbits of the exponential $`MD_4`$-groups. In this all section, we denote by $`G`$ one of the following groups: (with $`\phi \frac{\pi }{2}`$) $`G_{4,1,1};G_{4,1,2};G_{4,2,1(\lambda )};G_{4,2,2};G_{4,2,3(\phi )};G_{4,3,1(\lambda _1,\lambda _2)};G_{4,3,2(\lambda )};G_{4,3,3};G_{4,3,4(\lambda ,\phi )};G_{4,4,2}.`$ ### 3.1. Hamiltonian functions in canonical coordinates of $`\mathrm{\Omega }_F`$ Each element $`A𝔤`$ can be considered as a linear function $`\stackrel{~}{A}`$ on co-adjoint orbits $`(𝔤^{})`$: $`\stackrel{~}{A}(F^{}):=F^{},A,F^{}\mathrm{\Omega }_F`$. It is well-known that this function is the Hamiltonian function associated with the Hamiltonian vector field $`\xi _A`$, is defined on $`\mathrm{\Omega }_F`$ by $$(\xi _Af)(x):=\frac{d}{dt}f(x\mathrm{exp}(tA))|_{t=0},fC^{\mathrm{}}(\mathrm{\Omega }_F).$$ The Kirillov form $`\omega _F`$ is defined by the formula (17) $$\omega _F(\xi _A,\xi _B)=F,[A,B],A,B𝔤.$$ Denote by $`\psi `$ the indicated symplectomorphism from $`^2`$ onto $`\mathrm{\Omega }_F.`$ $$(p,q)^2\psi (p,q)\mathrm{\Omega }_F.$$ ###### Proposition 3.1.1. Each nontrivial orbit $`\mathrm{\Omega }_F𝔤^{}`$ of the co-adjoint representation of $`G`$, admits a global diffeomorphism $`\psi `$ $$\psi :(p,q)^2\psi (p,q)\mathrm{\Omega }_F$$ such that: * Hamiltonian function $`\stackrel{~}{A}=F^{},A,(F^{}\mathrm{\Omega }_F;A𝔤)`$ is of the form: $$\stackrel{~}{A}\psi (p,q)=\mathrm{\Phi }(q).p+\mathrm{\Psi }(q),$$ where $`\mathrm{\Phi }(q),\mathrm{\Psi }(q)`$ are $`C^{\mathrm{}}`$ -functions on $``$. * The Kirillov form (17) is (18) $$\omega =dpdq$$ Proof. * The diffeomorphism $`\psi `$ will be chosen case by case. 1. Case $`G_{4,1,1}`$ and $`\mathrm{\Omega }_F`$ defined by (1). We chose the diffeomorphism: $$\psi :^2\mathrm{\Omega }_F;(p,q)(q,\beta ,\gamma ,p).$$ Then, (19) $$\stackrel{~}{A}\psi (p,q)=dp+(aq+b\beta +c\gamma ).$$ 2. Case $`G_{4,1,2}`$ and $`\mathrm{\Omega }_F`$ defined by (2).Then we take $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha ,\beta ,\gamma e^q,p),$$ (20) $$\stackrel{~}{A}\psi (p,q)=dp+(c\gamma e^q+a\alpha +b\beta ).$$ 3. Case $`G_{4,2,1(\lambda )}`$ and $`\mathrm{\Omega }_F`$ defined by (3). $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha ,\beta e^{q\lambda },\gamma e^q,p),$$ (21) $$\stackrel{~}{A}\psi (p,q)=dp+(c\gamma e^q+a\alpha +b\beta e^{q\lambda }).$$ 4. Case $`G_{4,2,2}`$ and $`\mathrm{\Omega }_F`$ defined by (4). $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha ,\beta e^q,\beta qe^q+\gamma e^q,p),$$ (22) $$\stackrel{~}{A}\psi (p,q)=dp+c\beta qe^q+(b\beta +c\gamma )e^q+a\alpha .$$ 5. Case $`G_{4,2,3(\phi ),\phi \frac{\pi }{2}}`$ and $`\mathrm{\Omega }_F`$ defined by (5). $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha ,(\beta +i\gamma )e^{qe^{i\phi }},p),$$ (23) $$\stackrel{~}{A}\psi (p,q)=dp+(b+ic)(\beta +i\gamma )e^{qe^{i\phi }}+a\alpha .$$ 6. Case $`G_{4,3,1(\lambda _1,\lambda _2)}`$ and $`\mathrm{\Omega }_F`$ defined by (7). $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha e^{q\lambda _1},\beta e^{q\lambda _2},\gamma e^q,p),$$ (24) $$\stackrel{~}{A}\psi (p,q)=dp+a\alpha e^{q\lambda _1}+b\beta e^{q\lambda _2}+c\gamma e^q.$$ 7. Case $`G_{4,3,2(\lambda )}`$ and $`\mathrm{\Omega }_F`$ defined by (8). $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha e^{q\lambda },\alpha qe^{q\lambda }+\beta e^{q\lambda },\gamma e^q,p),$$ (25) $$\stackrel{~}{A}\psi (p,q)=dp+(a\alpha +bq\alpha +b\beta )e^{q\lambda }+c\gamma e^q.$$ 8. Case $`G_{4,3,3}`$ and $`\mathrm{\Omega }_F`$ defined by (9). $$\psi :^2\mathrm{\Omega }_F;(p,q)(\alpha e^q,\alpha qe^q+\beta e^q,\frac{1}{2}\alpha q^2e^q+\beta qe^q+\gamma e^q,p),$$ (26) $$\stackrel{~}{A}\psi (p,q)=dp+(a\alpha +b\alpha +b\beta +\frac{1}{2}c\alpha q^2+c\beta q+c\gamma )e^q.$$ 9. Case $`G_{4,3,4(\lambda ,\phi )},\phi \frac{\pi }{2}`$ and $`\mathrm{\Omega }_F`$ defined by (10). $$\psi :^2\mathrm{\Omega }_F;(p,q)((\alpha +i\beta )e^{qe^{i\phi }},\gamma e^{q\lambda },p),$$ (27) $$\stackrel{~}{A}\psi (p,q)=dp+(a+ib)(\alpha +i\beta )e^{qe^{i\phi }}+c\gamma e^{q\lambda }.$$ 10. Case $`G_{4,4,2}=_3`$ and 10.1 $`\mathrm{\Omega }_F`$ defined by (13). $$(p,q)^2\psi (p,q)=(\alpha e^q,0,0,p)\mathrm{\Omega }_F,$$ (28) $$\stackrel{~}{A}\psi (p,q)=dp+a\alpha e^q.$$ 10.2 $`\mathrm{\Omega }_F`$ defined by (14). $$(p,q)^2\psi (p,q)=(0,\beta e^q,0,p)\mathrm{\Omega }_F,$$ (29) $$\stackrel{~}{A}\psi (p,q)=dp+b\beta e^q.$$ 10.3 $`\mathrm{\Omega }_F`$ defined by (15). $$(p,q)^2\psi (p,q)=(\alpha e^q,\beta e^q,0,p)\mathrm{\Omega }_F,$$ (30) $$\stackrel{~}{A}\psi (p,q)=dp+a\alpha e^q+b\beta e^q.$$ 10.4 $`\mathrm{\Omega }_F`$ defined by (16) . $$(p,q)^2\psi (p,q)=(e^q,(\alpha \beta +\gamma p\gamma \delta )e^q,\gamma ,p)\mathrm{\Omega }_F,$$ (31) $$\stackrel{~}{A}\psi (p,q)=ae^q+b(\alpha \beta +\gamma p\gamma \delta )e^q+c\gamma +dp=$$ $$=(d+b\gamma e^q)p+ae^q+b(\alpha \beta \gamma \delta )e^q+c\gamma .$$ (see \[H3\] for more detail ). * We prove the Kirillov form on $`\mathrm{\Omega }_F`$ is $`\omega =dpdq`$ , namely for the case $`G_{4,2,3(\phi )},\phi \frac{\pi }{2}`$. From the Hamiltonian function $`\stackrel{~}{A}\psi (p,q)`$ we have $$\xi _A(f)=\{\stackrel{~}{A},f\}=d\frac{f}{q}(b+ic)(\beta +i\gamma )e^{i\phi }e^{qe^{i\phi }}\frac{f}{p},$$ with $`A=aX+bY+cZ+dT𝔤_{4,2,3(\phi )}`$ $$\xi _B(f)=\{\stackrel{~}{B},f\}=d^{}\frac{f}{q}(b^{}+ic^{})(\beta +i\gamma )e^{i\phi }e^{qe^{i\phi }}\frac{f}{p},$$ with $`B=a^{}X+b^{}Y+c^{}Z+d^{}T𝔤_{4,2,3(\phi )}`$. Thus, $$\xi _A\xi _B=dd^{}\frac{}{q}\frac{}{q}+(b+ic)(b^{}+ic^{})(\beta +i\gamma )^2e^{2i\phi }e^{2qe^{i\phi }}\frac{}{p}\frac{}{p}+$$ $$+[(db^{}d^{}b)+i(dc^{}d^{}c)](\beta +i\gamma )e^{i\phi }e^{qe^{i\phi }}\frac{}{p}\frac{}{q}$$ On the other hand, $`F^{},[A,B]=[(db^{}d^{}b)+i(dc^{}d^{}c)](\beta +i\gamma )e^{i\phi }e^{qe^{i\phi }}`$. This implies (18). The other cases are proved similarly. The proposition is hence completely proved. $`\mathrm{}`$ ###### DEFINITION 3.1.2. Each chart $`(\mathrm{\Omega }_F,\psi ^1)`$ satisfying 1. and 2. of the Proposition $`(\mathrm{3.1.1})`$ is called an adapted chart on $`\mathrm{\Omega }_F.`$ ### 3.2. Computation of operators $`\widehat{\mathrm{}}_A`$ Since $`\stackrel{~}{A}\psi (p,q)=\mathrm{\Phi }(q).p+\mathrm{\Psi }(q),`$ for $`A𝔤`$, one can prove that : $$P^r(\stackrel{~}{A},\stackrel{~}{B})=0r3,A,B𝔤.$$ From this we have the following proposition ###### Proposition 3.2.1. With $`A,B𝔤`$, the Moyal $``$-product satisfies the relation: (32) $$i\stackrel{~}{A}i\stackrel{~}{B}i\stackrel{~}{B}i\stackrel{~}{A}=i\stackrel{~}{[A,B]}$$ For each $`A𝔤`$ and the corresponding Hamiltonian function $`\stackrel{~}{A}`$, we denote by $`\mathrm{}_A`$ the operator acting on dense space $`L^2(^2,\frac{dpdq}{2\pi })`$ of smooth function by left $``$-multiplication by $`i\stackrel{~}{A}`$, i.e $`\mathrm{}_A(f)=i\stackrel{~}{A}f`$. The relation in the Proposition (3.2.1) gives us ###### Corollary 3.2.2. (33) $$\mathrm{}_{[A,B]}=\mathrm{}_A\mathrm{}_B\mathrm{}_B\mathrm{}_A:=[\mathrm{}_A,\mathrm{}_B]^{}$$ This implies that the corresponding $`A𝔤\mathrm{}_A=i\stackrel{~}{A}.`$ is a representation of the Lie algebra $`𝔤`$ on the space $`C^{\mathrm{}}(\mathrm{\Omega }_F)[[\frac{i}{2}]]`$ of formal power series. Let us denote by $`_p(f)`$ the partial Fourier transform ( define on $`Ł^2(^2,dpdq/2\pi )`$, for example ) of the function $`f`$ from the variable $`p`$ to the variable $`x`$ (see e.g \[MV\]), i.e $$_p(f)(x,q):=\frac{1}{\sqrt{2\pi }}_{}e^{ipx}f(p,q)𝑑p$$ and $`_p^1(f)(p,q)`$ the inverse Fourier transform. We have following obvious identities: ###### Lemma 3.2.3. With $`f,\phi Ł^2(^2,dpdq/2\pi )`$ * $`_p_p^1(f)=i_p^1(x.f)`$ * $`_p(p.\phi )=i_x_p(\phi )`$ * $`P^r(\stackrel{~}{A},_p^1(f))=(1)^r_q^r(\mathrm{\Psi })_p^r_p^1(f)`$$`r2.`$ Now we denote $`\widehat{\mathrm{}}_A:=_p\mathrm{}_A_p^1`$ with $`A𝔤.`$ ###### DEFINITION 3.2.4. Let $`\mathrm{\Omega }_F`$ be K-orbit of co-adjoint representations of Lie group $`G`$. With $`A`$ running over the Lie algebra $`𝔤=Lie(G),`$ $`(\mathrm{\Omega }_F,\widehat{\mathrm{}}_A)`$ is called quantum co-adjoint orbits of Lie group $`G`$. ###### Theorem 3.2.5. For each $`A𝔤`$ and for each compactly supported $`C^{\mathrm{}}`$-function $`fC_c^{\mathrm{}}(^2)`$, we have $$\widehat{\mathrm{}}_A(f)=\mathrm{\Phi }(q\frac{x}{2})(\frac{1}{2}_q_x)f+i\mathrm{\Psi }(q\frac{x}{2})f$$ and setting new variables $`s=q\frac{x}{2},t=q+\frac{x}{2}`$, then (34) $$\widehat{\mathrm{}}_A(f)=\mathrm{\Phi }(s)\frac{f}{s}+i\mathrm{\Psi }(s)f|_{(s,t)}\text{i.e:}\widehat{\mathrm{}}_A=[\mathrm{\Phi }(s)\frac{}{s}+i\mathrm{\Psi }(s)]|_{(s,t)}$$ Proof. It is easy to see that: $$P^0(\stackrel{~}{A},_p^1(f))=[\mathrm{\Phi }(q).p+\mathrm{\Psi }(q)]_p^1(f)$$ $$P^1(\stackrel{~}{A},_p^1(f))=\mathrm{\Phi }(q)_q_p^1(f)[p_q\mathrm{\Phi }(q)+_q\mathrm{\Psi }(q)]_p^1(f)$$ $$P^r(\stackrel{~}{A},_p^1(f))=(1)^r_q^r\mathrm{\Psi }_p^r_p^1(f)r2.$$ From this and lemma (3.2.3), we have: $$\widehat{\mathrm{}}_A(f)=_p\mathrm{}_A_p^1(f)=i_p(\stackrel{~}{A}_p^1(f))=i_p\left(\underset{r0}{}\left(\frac{1}{2i}\right)^r\frac{1}{r!}P^r(\stackrel{~}{A},_p^1(f))\right)=$$ $$=i_p\{[\mathrm{\Phi }(q).p+\mathrm{\Psi }(q)]_p^1(f)+\frac{1}{1!}\frac{1}{2i}(\mathrm{\Phi }(q)_q_p^1(f)[p_q\mathrm{\Phi }(q)+_q\mathrm{\Psi }(q)]_p^1(f))+\mathrm{}+$$ $$+\frac{1}{r!}(\frac{1}{2i})^r(1)^r_q^r\mathrm{\Psi }_p^r_p^1(f))+\mathrm{}\}=\mathrm{\Phi }(q\frac{x}{2})(\frac{1}{2}_q_x)f+i\mathrm{\Psi }(q\frac{x}{2})f$$ Theorem is proveed. $`\mathrm{}`$ As a direct consequence of symbol $`\widehat{\mathrm{}}_A,`$ we have ###### Corollary 3.2.6. $`A,B𝔤,`$ $$\widehat{\mathrm{}}_A\widehat{\mathrm{}}_B\widehat{\mathrm{}}_B\widehat{\mathrm{}}_A=\widehat{\mathrm{}}_{[A,B]}$$ From theorems 2.0.3 and 3.2.5 we obtained the quantum half planes ,the quantum planes, the quantum hyperbolic cylinders, quantum hyperbolic paraboloids …of the corresponding groups. At the same time, we have also unitary representations of these groups : $$T(expA)=\mathrm{exp}(\widehat{\mathrm{}}_A)=\mathrm{exp}\left([\mathrm{\Phi }(s)\frac{}{s}+i\mathrm{\Psi }(s)]|_{(s,t)}\right).$$ ## 4. The case of groups $`G_{4,2,3(\frac{\pi }{2})},G_{4,2,4},G_{4,3,4(\frac{\pi }{2})},G_{4,4,1}`$. ### 4.1. The local diffeomorphisms For the Lie group of affine transformations of the complex straight line $`G_{4,2,4}=Aff()`$ (see \[DH2\]), we replaced the global diffeomorphism $`\psi `$ by a local diffeomorphism and obtained ###### Proposition 4.1.1. Fixing the local diffeomorphism $`\psi _k(k)`$ $$\psi _k:\times _k\times _k$$ $$(z,w)(z,e^w),$$ we have 1. For any element $`Aaff()`$, the corresponding Hamiltonian function $`\stackrel{~}{A}`$ in local coordinates $`(z,w)`$ of the orbit $`\mathrm{\Omega }_F`$ is of the form $$\stackrel{~}{A}\psi _k(z,w)=\frac{1}{2}[\alpha z+\beta e^w+\overline{\alpha }\overline{z}+\overline{\beta }e^{\overline{w}}]$$ 2. In local coordinates $`(z,w)`$ of the orbit $`\mathrm{\Omega }_F`$, the Kirillov form $`\omega `$ is of the form $$\omega =\frac{1}{2}[dzdw+d\overline{z}d\overline{w}].$$ Analogously, for the groups $`G_{4,2,3(\phi )},G_{4,3,4(\phi )}`$ with $`\phi =\frac{\pi }{2}`$, we also replace the global diffeomorphism $`\psi `$ by a local diffeomorphism $`\psi _k(k).`$ * For $`G=G_{4,2,3(\frac{\pi }{2})};\mathrm{\Omega }_F=\{(\alpha ,(\beta +i\gamma )e^{is},t)|s,t\}`$, $$\psi _k:\times (2k\pi ,2\pi +2k\pi )\mathrm{\Omega }_F.$$ $$(p,q)(\alpha ,(\beta +i\gamma )e^{iq},p)$$ Then the corresponding Hamiltonian fuction is $`\stackrel{~}{A}\psi _k(p,q)=dp+(b+ic)(\beta +i\gamma )e^{iq}+a\alpha .`$ * For $`G=G_{4,3,4(\frac{\pi }{2})};\mathrm{\Omega }_F=\{((\alpha +i\beta )e^{is},\gamma e^{\lambda s},t)|s,t\}`$, $$\psi _k:\times (2k\pi ,2\pi +2k\pi )\mathrm{\Omega }_F.$$ $$(p,q)((\alpha +i\beta )e^{iq},\gamma e^{q\lambda },p)$$ We have $`\stackrel{~}{A}\psi _k(p,q)=dp+(a+ib)(\alpha +i\beta )e^{iq}+c\gamma e^{q\lambda }.`$ At last, for the Lie group $`G_{4,4,1}=_j_3,`$ which is not exponential group, we obtain ###### Proposition 4.1.2. Let us denote $`𝐈_k=(2k\pi ,2\pi +2k\pi ),k`$. Each non-trivial orbit $`\mathrm{\Omega }_F`$ (in $`𝔤^{}`$) of co-adjoint representation of $`G_{4,4,1}`$ admits local charts $`(\times 𝐈_k,\psi _k^1)`$ or $`(_\pm \times 𝐈_k,\psi _k^1)`$ such that: 1. If $`\mathrm{\Omega }_F`$ is defined by (11) and $`A𝔤_{4,4,1},`$ then $$\stackrel{~}{A}\psi _k(p,q)=dp+\frac{1}{2}[a(\alpha +i\beta )+b(\beta i\alpha )]e^{iq}+\frac{1}{2}[a(\alpha i\beta )+b(\beta +i\alpha )]e^{iq}$$ (35) $$=dp+(a\alpha +b\beta )cosq+(b\alpha a\beta )sinq$$ and the Kirillov form then is $`\omega =dpdq.`$ 2. If $`\mathrm{\Omega }_F`$ is defined by (12) and $`A𝔤_{4,4,1},`$ then $$\stackrel{~}{A}\psi _k(p,q)=\frac{d}{2\gamma }p^2+[\frac{a}{2}(e^{iq}+e^{iq})+\frac{b}{2i}(e^{iq}e^{iq})]p+c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }$$ (36) $$=\frac{d}{2\gamma }p^2+(acosq+bsinq)p+c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }$$ and we have the Kirillov form to be $`\omega =\frac{\gamma }{p}dpdq.`$ Proof. 1. We consider the following diffeomorphism: $$\psi _k:\times 𝐈_k\mathrm{\Omega }_F$$ $$(p,q)(\frac{1}{2}(\alpha +i\beta )e^{iq}+\frac{1}{2}(\alpha i\beta )e^{iq};\frac{1}{2}(\beta i\alpha )e^{iq}+\frac{1}{2}(\beta +i\alpha )e^{iq};0;p)$$ With each $`F^{}\mathrm{\Omega }_F`$, $$F^{}=[\frac{1}{2}(\alpha +i\beta )e^{iq}+\frac{1}{2}(\alpha i\beta )e^{iq}]X^{}+[\frac{1}{2}(\beta i\alpha )e^{iq}+\frac{1}{2}(\beta +i\alpha )e^{iq}]Y^{}+pT^{}$$ and $`A=aX+bY+cZ+dT𝔤_{4,4,1},`$ we have $$\stackrel{~}{A}(F^{})=F^{},A=\frac{a}{2}[(\alpha +i\beta )e^{iq}+(\alpha i\beta )e^{iq}]+\frac{b}{2}[(\beta i\alpha )e^{iq}+(\beta +i\alpha )e^{iq}]+dp.$$ It follows that $$\xi _A(f)=d\frac{f}{q}\frac{i}{2}[a(\alpha +i\beta )+b(\beta i\alpha )]e^{iq}+\frac{i}{2}[a(\alpha i\beta )+b(\beta +i\alpha )]e^{iq}\frac{f}{p}.$$ By analogy, $$\xi _B(f)=d^{}\frac{f}{q}\frac{i}{2}[a^{}(\alpha +i\beta )+b^{}(\beta i\alpha )]e^{iq}+\frac{i}{2}[a^{}(\alpha i\beta )+b^{}(\beta +i\alpha )]e^{iq}\frac{f}{p}.$$ Thus, (37) $$\xi _A\xi _B=dd^{}\frac{}{q}\frac{}{q}\frac{1}{4}\{[a^{}(\alpha +i\beta )+b^{}(\beta i\alpha )]e^{iq}[a^{}(\alpha i\beta )+b^{}(\beta +i\alpha )]e^{iq}\}$$ $$\times \{[a(\alpha +i\beta )+b(\beta i\alpha )]e^{iq}[a(\alpha i\beta )+b(\beta +i\alpha )]e^{iq}\}\frac{}{p}\frac{}{p}+\frac{1}{2}\{[(db^{}d^{}b)(\alpha +i\beta )+$$ $$+(ad^{}a^{}d)(\beta i\alpha )]e^{iq}+[(db^{}d^{}b)(\alpha i\beta )+(ad^{}a^{}d)(\beta +i\alpha )]e^{iq}\}\frac{}{p}\frac{}{q}$$ On the other hand, $`[A,B]=(db^{}d^{}b)X+(ad^{}a^{}d)Y+(ab^{}a^{}b)Z`$ implies (38) $$F^{},[A,B]=\frac{1}{2}[(db^{}d^{}b)(\alpha +i\beta )+(ad^{}a^{}d)(\beta i\alpha )]e^{iq}+$$ $$+\frac{1}{2}[(db^{}d^{}b)(\alpha i\beta )+(ad^{}a^{}d)(\beta +i\alpha )]e^{iq}=\frac{1}{2}\{[(db^{}d^{}b)(\alpha +i\beta )+$$ $$+(ad^{}a^{}d)(\beta i\alpha )]e^{iq}+[(db^{}d^{}b)(\alpha i\beta )+(ad^{}a^{}d)(\beta +i\alpha )]e^{iq}$$ (37) and (38) imply that the Kirillov form is $`\omega =dpdq`$. 2.For the case $`\gamma 0`$ we chose $$\psi _k:(_+\times 𝐈_k)\mathrm{\Omega }_F$$ $$(p,q)(pcosq,psinq,\gamma ,\frac{1}{2\gamma }(p^2+2\gamma \delta \alpha ^2\beta ^2))$$ $$(\text{or}\psi _k:(_{}\times 𝐈_k)\mathrm{\Omega }_F).$$ Then, $`F^{}\mathrm{\Omega }_F,F^{}=pcosqX^{}+psinqY^{}+\gamma Z^{}+\frac{1}{2\gamma }(p^2+2\gamma \delta \alpha ^2\beta ^2)T^{}`$ and $`A=aX+bY+cZ+dT𝔤_{4,4,1}`$, we have $$\stackrel{~}{A}\psi _{(U,k)}=apcosq+bpsinq+c\gamma +\frac{d}{2\gamma }(p^2+2\gamma \delta \alpha ^2\beta ^2)=$$ $$=\frac{d}{2\gamma }p^2+[\frac{a}{2}(e^{iq}+e^{iq})+\frac{b}{2i}(e^{iq}e^{iq})]p+c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }.$$ It follows that $$\xi _A(f)=[\frac{d}{\gamma }p+\frac{1}{2i}((ai+b)e^{iq}+(aib)e^{iq})]\frac{f}{q}\frac{p}{2}[(ai+b)e^{iq}(aib)e^{iq}]\frac{f}{p}.$$ By analogy, $$\xi _B(f)=[\frac{d^{}}{\gamma }p+\frac{1}{2i}((a^{}i+b^{})e^{iq}+(a^{}ib^{})e^{iq})]\frac{f}{q}\frac{p}{2}[(a^{}i+b^{})e^{iq}(a^{}ib^{})e^{iq}]\frac{f}{p}.$$ Thus, (39) $$\xi _A\xi _B=\frac{p}{2}[(ai+b)e^{iq}(aib)e^{iq}\frac{p}{2}[(a^{}i+b^{})e^{iq}(a^{}ib^{})e^{iq}]\frac{}{p}\frac{}{p}+$$ $$+[\frac{d}{\gamma }p+\frac{1}{2i}((ai+b)e^{iq}+(aib)e^{iq})][\frac{d^{}}{\gamma }p+\frac{1}{2i}((a^{}i+b^{})e^{iq}+(a^{}ib^{})e^{iq})]\frac{}{q}\frac{}{q}+$$ $$+\left\{\frac{p^2}{2\gamma }\left[\left((da^{}d^{}a)i+(db^{}d^{}b)\right)e^{iq}\left((da^{}d^{}a)i(db^{}d^{}b)\right)e^{iq}\right]+p(ab^{}a^{}b)\right\}\frac{}{p}\frac{}{q}.$$ On the other hands, (40) $$F^{},[A,B]=(db^{}d^{}b)pcosq+(ad^{}a^{}d)psinq+(db^{}d^{}b)\gamma =$$ $$=\frac{p}{2}[((a^{}dd^{}a)i+(db^{}d^{}b))e^{iq}((da^{}d^{}a)i(db^{}d^{}b))e^{iq}]+\gamma (ab^{}a^{}b).$$ From (39) and (40) we see that (17) is of the form $`\omega =\frac{\gamma }{p}dpdq.`$ $`\mathrm{}`$ ### 4.2. Computation of operators $`\widehat{\mathrm{}}_A^{(k)}`$ It is easy to prove that, * If $`G=G_{4,2,3(\frac{\pi }{2})}`$ then $$\widehat{\mathrm{}}_A^{(k)}=\left(d\frac{}{_s}+i(b+ic)(\beta +i\gamma )e^{is}+a\alpha \right)|_{(s,t)}$$ * If $`G=G_{4,3,4(\frac{\pi }{2})}`$ then $$\widehat{\mathrm{}}_A^{(k)}=\left(d\frac{}{_s}+i(a+ib)(\alpha +i\beta )e^{is}+c\gamma e^{s\lambda }\right)|_{(s,t)}$$ In \[DH2\], we proved the following result for $`G_{4,2,4}`$ Let $`_z`$(f) denote the partial Fourier transform of the function $`f`$ from the variable $`z=p_1+ip_2`$ to the variable $`\xi =\xi _1+i\xi _2`$, i.e: $$_z(f)(\xi ,w)=\frac{1}{2\pi }_{R^2}e^{iRe(\xi \overline{z})}f(z,w)𝑑p_1𝑑p_2$$ and $$_z^1(f)(z,w)=\frac{1}{2\pi }_{R^2}e^{iRe(\xi \overline{z})}f(\xi ,w)𝑑\xi _1𝑑\xi _2$$ the inverse Fourier transform. ###### Theorem 4.2.1. ( see \[DH2\],Proposition 3.4) For each $`A=\left(\begin{array}{cc}\alpha & \beta \\ 0& 0\end{array}\right)aff()`$ and for each compactly supported $`C^{\mathrm{}}`$-function $`fC_c^{\mathrm{}}(\times _k)`$, we have: (41) $$\widehat{\mathrm{}}_A^{(k)}(f):=_z\mathrm{}_A^{(k)}_z^1(f)=$$ $$=[\alpha (\frac{1}{2}_w_{\overline{\xi }})f+\overline{\alpha }(\frac{1}{2}_{\overline{w}}_\xi )f+\frac{i}{2}(\beta e^{w\frac{1}{2}\overline{\xi }}+\overline{\beta }e^{\overline{w}\frac{1}{2}\xi })f]$$ i.e $$\widehat{\mathrm{}}_A^{(k)}=\alpha \frac{}{u}+\overline{\alpha }\frac{}{\overline{u}}+\frac{i}{2}(\beta e^u+\overline{\beta }e^{\overline{u}});u=w\frac{1}{2}\overline{\xi };v=w+\frac{1}{2}\overline{\xi }.$$ Now we consider group $`G_{4,4,1}`$ : ###### Theorem 4.2.2. For each $`A𝔤_{4,4,1}`$ and for each compactly supported $`C^{\mathrm{}}`$-function $`fC_0^{\mathrm{}}(\times 𝐈_k)`$, we have: 1. If $`\stackrel{~}{A}`$ is defined by (35 ) then $$\widehat{\mathrm{}}_A^{(k)}(f)=\left(d_sf+i[(a\alpha +b\beta )coss+(b\alpha a\beta )sins]f\right)|_{(s,t)}$$ 2. If $`\stackrel{~}{A}`$ is defined by (36 ) then $$\widehat{\mathrm{}}_A^{(k)}(f)=i.([\frac{d}{2\gamma }(i_x+\frac{1}{2\gamma }_x_q)^2]f+\mathrm{\Gamma }.f)|_{(x,q)}$$ $$\frac{1}{2}\left(_x[(abi)\mathrm{\Delta }(f)+(a+bi)\mathrm{\Delta }^1(f)]+\frac{i}{2\gamma }_x[(abi)\mathrm{\Theta }(f)+(a+bi)\mathrm{\Theta }^1(f)]\right)|_{(x,q)}$$ where $$\mathrm{\Gamma }=c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }$$ $$\mathrm{\Delta }(f)=\mathrm{exp}[iq+_x((\frac{x}{2\gamma }).f)]=e^{iq}.\underset{r=0}{}\frac{1}{r!}\frac{^r}{_{x^r}}((\frac{x}{2\gamma })^r.f)$$ $$\mathrm{\Theta }(f)=\mathrm{exp}[iq+_x((\frac{x}{2\gamma })._qf)]=e^{iq}.\underset{r=0}{}\frac{1}{r!}\frac{^r}{_{x^r}}((\frac{x}{2\gamma })^r._qf).$$ To prove the theorem, we need the following obvious lemma, which is a direct conseqnence of the definition of $`_p`$ and $`_p^1`$. ###### Lemma 4.2.3. With $`r1`$ * $`_{p^r}^r\left(_p^1(f)\right)=i^r_p^1(x^r.f)`$ * $`_p\left(p^r_p^1(f)\right)=i^r_{x^r}^r(f)`$ * $`_p\left(p^r_{p^{r1}}^{r1}_p^1(f)\right)=i^{2r1}_{x^r}^r(x^{r1}.f).`$ $`\mathrm{}`$ Proof. First case is proved like as Theorem 3.2.5. We prove only the second case. One can write (36) as $$\stackrel{~}{A}\psi (p,q)=\frac{d}{2\gamma }p^2+\frac{p}{2}[(abi)e^{iq}+(a+bi)e^{iq}]+c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }$$ and remark that in the coordinates $`(p,q)`$ correspond to the form $`\omega =\frac{\gamma }{p}dpdq`$, $$\mathrm{\Lambda }^1=\left(\begin{array}{cc}0& \frac{p}{\gamma }\\ \frac{p}{\gamma }& 0\end{array}\right).$$ Denoting $`_p^1(f)=v,`$ we have $$P^0=\left\{\frac{d}{2\gamma }p^2+\frac{p}{2}[(abi)e^{iq}+(a+bi)e^{iq}]+c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }\right\}v$$ $$P^1=[\frac{dp^2}{\gamma ^2}+\frac{p}{2\gamma }((abi)e^{iq}+(a+bi)e^{iq})]_qv\frac{ip^2}{2\gamma }[(abi)e^{iq}(a+bi)e^{iq}]_pv$$ $$P^2=\frac{1}{2\gamma ^2}\{i^2[(abi)e^{iq}+(a+bi)e^{iq}]p^3_{p^2}^2v2i[(abi)e^{iq}(a+bi)e^{iq}]p^2_{pq}^2v\}+\frac{dp^2}{\gamma ^3}_{q^2}^2v$$ $$P^3=\frac{1}{2\gamma ^3}\{(i)^3[(abi)e^{iq}(a+bi)e^{iq}]p^4_{p^3}^3v+3(i)^2[(abi)e^{iq}+(a+bi)e^{iq}]p^3_{p^2q}^3v\}$$ $$P^4=\frac{1}{2\gamma ^4}\{(i)^4[(abi)e^{iq}+(a+bi)e^{iq}]p^5_{p^4}^4v+4(i)^3[(abi)e^{iq}(a+bi)e^{iq}]p^4_{p^3q}^4v\}$$ By analogy, for $`r4`$ we have $$P^r=\frac{1}{2\gamma ^r}\left((i)^r[(abi)e^{iq}+(1)^r(a+bi)e^{iq}]p^{r+1}_{p^r}^rv\right)+$$ $$+\frac{r(i)^{r1}}{2\gamma ^r}\left([(abi)e^{iq}+(1)^{r1}(a+bi)e^{iq}]p^r_{p^{r1}q}^rv\right).$$ As $`\stackrel{~}{A}`$ is defined by (36), we obtain: (42) $$\widehat{\mathrm{}}_A^{(k)}(f):=_p\mathrm{}_A_p^1(f)=_p(i\stackrel{~}{A}_p^1(f))=i\underset{r0}{}\left(\frac{1}{2i}\right)^r_p(P^r(\stackrel{~}{A},_p^1(f))=$$ $$=i(c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma })_pv+i\left[_p(\frac{d}{2\gamma }p^2v)+\frac{1}{1!}\frac{1}{2i}_p(\frac{d}{\gamma ^2}p^2_qv)+\frac{1}{2!}\left(\frac{1}{2i}\right)^2_p(\frac{d}{\gamma ^3}p^2_{q^2}^2v)\right]+$$ $$+\underset{r0}{}\frac{1}{r!}\left(\frac{1}{2i}\right)^r\frac{1}{2\gamma ^r}\{(i)^r[(abi)e^{iq}+(1)^r(a+bi)e^{iq}]_p(p^{r+1}_{p^r}^rv)+r(i)^{r1}[(abi)e^{iq}+$$ $$+(1)^{r1}(a+bi)e^{iq}]_p(p^r_{p^{r1}q}^rv)\}=i(c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma })f+i\frac{d}{2\gamma }(i_x+\frac{1}{2\gamma }_x_q)^2f+$$ $$+\frac{i}{2}\underset{r0}{}\frac{1}{r!}\left(\frac{1}{2\gamma }\right)^r(abi)e^{iq}_p(p^{r+1}_{p^r}^rv)+\frac{i}{2}\underset{r0}{}\frac{1}{r!}\left(\frac{1}{2\gamma }\right)^r(a+bi)e^{iq}_p(p^{r+1}_{p^r}^rv)+$$ $$+\frac{i}{2}\frac{1}{\gamma }\underset{r1}{}\frac{1}{(r1)!}\left(\frac{1}{2\gamma }\right)^{r1}\frac{1}{2i}(abi)e^{iq}_p(p^r_{p^{(r1)q}}^rv)+\frac{i}{2}\frac{1}{\gamma }\underset{r1}{}\frac{1}{(r1)!}\left(\frac{1}{2\gamma }\right)^{r1}\frac{1}{2i}(a+bi)$$ $$\times e^{iq}_p(p^r_{p^{(r1)q}}^rv)=i[c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }+i\frac{d}{2\gamma }(i_x+\frac{1}{2\gamma }_x_q)^2]f+\frac{i}{2}(abi)e^{iq}$$ $$\times \underset{r0}{}\frac{1}{r!}\left(\frac{1}{2\gamma }\right)^ri^{2r+1}_{x^{r+1}}^{r+1}(x^r.f)+\frac{i}{2}(a+bi)e^{iq}\underset{r0}{}\frac{1}{r!}\left(\frac{1}{2\gamma }\right)^ri^{2r+1}_{x^{r+1}}^{r+1}(x^r.f)+$$ $$+\frac{1}{4\gamma }(abi)e^{iq}\underset{r1}{}\frac{1}{(r1)!}\left(\frac{1}{2\gamma }\right)^{r1}i^{2r1}_{p^{r1}}^{r1}(x^{r1}._qf)+$$ $$+\frac{1}{4\gamma }(a+bi)e^{iq}\underset{r1}{}\frac{1}{(r1)!}\left(\frac{1}{2\gamma }\right)^{r1}i^{2r1}_{p^{r1}}^{r1}(x^{r1}._qf)=$$ $$=i[c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }+i\frac{d}{2\gamma }(i_x+\frac{1}{2\gamma }_x_q)^2]f\frac{1}{2}_x[(abi)e^{iq}.\underset{r0}{}\frac{1}{r!}\left(\frac{1}{2\gamma }\right)^r_{x^r}^r(x^r.f)+$$ $$+(a+bi)e^{iq}.\underset{r0}{}\frac{1}{r!}\left(\frac{1}{2\gamma }\right)^r_{x^r}^r(x^r.f)+\frac{i}{2\gamma }(abi)e^{iq}.\underset{r1}{}\frac{1}{(r1)!}\left(\frac{1}{2\gamma }\right)^{r1}_{x^{r1}}^{r1}(x^{r1}._qf)+$$ $$+\frac{i}{2\gamma }(a+bi)e^{iq}.\underset{r1}{}\frac{1}{(r1)!}\left(\frac{1}{2\gamma }\right)^{r1}_{x^{r1}}^{r1}(x^{r1}._qf)]=$$ $$=i[c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }+i\frac{d}{2\gamma }(i_x+\frac{1}{2\gamma }_x_q)^2]f$$ $$\frac{1}{2}_x\left((abi)e^{iq}e^{_x((\frac{x}{2\gamma }).f)}+(a+bi)e^{iq}e^{_x((\frac{x}{2\gamma }).f)}\right)$$ $$\frac{i}{4\gamma }_x\left((abi)e^{iq}e^{_x((\frac{x}{2\gamma })._qf)}+(a+bi)e^{iq}e^{_x((\frac{x}{2\gamma })._qf)}\right)=$$ $$=i[c\gamma +d\delta d\frac{\alpha ^2+\beta ^2}{2\gamma }+\frac{d}{2\gamma }(i_x+\frac{1}{2\gamma }_x_q)^2]f$$ $$\frac{1}{2}_x((abi)\mathrm{exp}[iq+_x((\frac{x}{2\gamma }).f)]+(a+bi)\mathrm{exp}[iq+_x((\frac{x}{2\gamma }).f)])$$ $$\frac{i}{4\gamma }_x((abi)\mathrm{exp}[iq+_x((\frac{x}{2\gamma })._qf)]+(a+bi)\mathrm{exp}[iq+_x((\frac{x}{2\gamma })._qf)])$$ The theorem is therefore completely proved. $`\mathrm{}`$ Thus, we obtained all the operators $`\widehat{\mathrm{}}_A;\widehat{\mathrm{}}_A^k`$, which provides (global or local) representations of the $`MD_4`$-algebras. At last, as $`\widehat{\mathrm{}}_A,\widehat{\mathrm{}}_A^k`$ are representations of the $`MD_4`$-algebras, we have operators : $`\mathrm{exp}(\widehat{\mathrm{}}_A);\mathrm{exp}(\widehat{\mathrm{}}_A^k)`$ are representations of the corresponding connected and simply connected $`MD_4`$-groups. We say that they are the representations of $`MD_4`$-groups arising from the reduction of the procedure of deformation quantization. ACKNOWLEDGMENT The author would like to acknowledge the support of Seminar ”Deformation Quantization and Applications” under the supervision by Professor Do Ngoc Diep and would like to thank Dr. Nguyen Viet Dung for his useful comments .
warning/0003/nucl-ex0003002.html
ar5iv
text
# Quasi-free Compton Scattering from the Deuteron and Nucleon Polarizabilities ## Abstract Cross sections for quasi-free Compton scattering from the deuteron were measured for incident energies of E<sub>γ</sub>=236–260 MeV at the laboratory angle $`\theta _\gamma ^{}=135^{}`$. The recoil nucleons were detected in a liquid-scintillator array situated at $`\theta _N=20^{}`$. The measured differential cross sections were used, with the calculations of Levchuk et al., to determine the polarizabilities of the bound nucleons. For the bound proton, the extracted values were consistent with the accepted value for the free proton. Combining our results for the bound neutron with those from Rose et al., we obtain one-sigma constraints of $`\overline{\alpha }_n=7.614.0`$ and $`\overline{\beta }_n=1.27.6`$. The electric $`\overline{\alpha }`$ and magnetic $`\overline{\beta }`$ polarizabilities constitute the first-order responses of the internal structure of a nucleon to externally applied electric and magnetic fields. Compton scattering from the proton has been used extensively to determine the polarizabilities of the proton (see Ref. and references contained therein). Such measurements are sensitive to the sum (difference) of the polarizabilities for photons scattered at forward (backward) angles. Ref. reports the experimental status of the proton polarizabilities, in the usual units of $`10^4`$ fm<sup>3</sup>: $`(\overline{\alpha }\overline{\beta })_p`$ $`=`$ $`10.0\pm 1.5\pm 0.9,`$ (1) $`(\overline{\alpha }+\overline{\beta })_p`$ $`=`$ $`15.2\pm 2.6\pm 0.2,`$ (2) where the first error is the combined statistical and systematic, and the second is due to the model dependence of the dispersion-relation extraction method. Disperson sum rules relate the sum of the polarizabilities to the nucleon photoabsorption cross section. The generally accepted results were $`(\overline{\alpha }+\overline{\beta })_p`$ $`=`$ $`14.2\pm 0.5,`$ (3) $`(\overline{\alpha }+\overline{\beta })_n`$ $`=`$ $`15.8\pm 0.5.`$ (4) There have been recent reevaluations of these sum rules which yield $`13.69\pm 0.14`$ ($`14.40\pm 0.66`$) and $`14.0\pm 0.5`$ ($`15.2\pm 0.5`$) for the proton (neutron). The proton polarizabilities, obtained from Eqs. 1 and 3, are $$\overline{\alpha }_p=12.1\pm 0.8\pm 0.5,\overline{\beta }_p=2.10.80.5.$$ (5) The status of the neutron polarizabilities is much less satisfactory (see Ref. and references contained therein). Measurements of the electric polarizability of the neutron have been done by low-energy neutron scattering from the Coulomb field of a heavy nucleus. The extracted values fall in the range $`\alpha _n`$=0–19 . Elastic Compton scattering from the deuteron has also been used to extract information on the nucleon polarizabilities. Using their theoretical model, Levchuk and L’vov have reported values of $$(\overline{\alpha }\overline{\beta })_n=2\pm 3,(\overline{\alpha }+\overline{\beta })_n=20\pm 3,$$ (6) from fitting to the data of Refs. . The large discrepancy between $`(\overline{\alpha }\overline{\beta })_n`$ and $`(\overline{\alpha }\overline{\beta })_p`$ (Eq. 1) may be indicative of shortcomings in the theoretical models used to extract the polarizabilities from the $`d(\gamma ,\gamma )d`$ reaction. The quasi-free Compton scattering reaction $`d(\gamma ,\gamma ^{}n)p`$, in which the scattered photon is detected in coincidence with the recoil neutron, can be used to minimize the model dependence of the extracted polarizabilities. There has been one measurement reported on this reaction using bremsstrahlung photons with an endpoint of 130 MeV . A most-probable value of $`\overline{\alpha }_n=10.7`$ was obtained with an upper limit of 14.0 but with no constraint on the lower limit. In addition, the model dependence is strong at these lower energies. Levchuk et al. have determined that, within the context of their model, the sensitivity to the neutron amplitude is maximized (and model dependence minimized) for E<sub>γ</sub>=200–300 MeV and backward angles for the scattered photons. For $`E_\gamma `$=247 MeV and $`\theta _\gamma ^{}`$=$`135^{}`$, the contribution of the spectator nucleon modifies the cross section by only $``$4% at the quasi-free peak. Furthermore, the free-neutron cross section can be related to the cross section in the center of the neutron quasi-free peak via a spectator formula $$\frac{d\sigma (\gamma n\gamma ^{}n)}{d\mathrm{\Omega }_\gamma ^{}}=\frac{(2\pi )^3}{u^2(0)}\frac{E_\gamma E_\gamma ^{}}{|𝐩_n|mE_\gamma ^{}^{(n)2}}\frac{d^3\sigma (\gamma d\gamma ^{}np)}{d\mathrm{\Omega }_\gamma ^{}d\mathrm{\Omega }_ndE_n},$$ (7) where $`u(0)`$ is the S-wave amplitude of the deuteron wave function at zero momentum, and $`E_\gamma ^{}^{(n)}`$ is the photon energy in the rest frame of the final neutron. The cross section on the right side of the equation must be corrected for the small contribution of the spectator proton (its pole diagram and final state interactions). The present measurement was performed at the Saskatchewan Accelerator Laboratory (SAL). The $`d(\gamma ,\gamma ^{}n)p`$ and $`d(\gamma ,\gamma ^{}p)n`$ cross sections were measured simultaneously, in kinematics that emphasized the quasi-free reaction. Data were also obtained from the free proton, using the same apparatus, for calibration and normalization purposes. An electron beam of 292 MeV and $``$60% duty factor produced bremsstrahlung photons, which were tagged in the energy range 236–260 MeV, with a resolution of 0.4 MeV. The average, integrated tagged flux was $`6\times 10^6`$ photons/s with a tagging efficiency of $``$45%. The total integrated luminosity for the deuterium measurements was $`1.25\times 10^{37}`$ photons/cm<sup>2</sup>. The cryogenic target cell was a vertical cylinder (12 cm in diameter), with Mylar walls, containing liquid hydrogen ($`p(\gamma ,\gamma )`$) or liquid deuterium ($`d(\gamma ,\gamma ^{}N`$) and $`d(\gamma ,pn)`$). The quasi-free scattering of 247 MeV photons to $`135^{}`$ from deuterium results in the recoil nucleon being confined to a forward cone centered around 18. These nucleons have a central kinetic energy of $``$77 MeV and were detected in a liquid-scintillator array situated at 20, subtending a solid angle of approximately 0.11 sr. The array consisted of 85 Lucite-walled cells, filled with BC-505 liquid scintillator. Between the cells and the target were thin plastic scintillators which acted as veto detectors for neutrons or as $`\mathrm{\Delta }`$E detectors for protons. The neutron efficiency of the array was measured at the beginning of the experiment via the $`d(\gamma ,pn)`$ reaction. A plastic-scintillator $`\mathrm{\Delta }`$E$``$E telescope was placed at 80 to detect the protons while the neutron array was located at $`77^{}`$. The detection efficiency was measured for neutrons with kinetic energies of 50–100 MeV, covering the range of interest for the quasi-free reaction. With the applied threshold of 11 MeV (6 MeVelectron equivalent) the efficiency ranged from 5.2–5.9%. Photons scattered from the target were detected in the large-volume Boston University NaI (BUNI) gamma-ray spectrometer . BUNI is composed of five optically-isolated segments of NaI, the core and four surrounding quadrants. Plastic-scintillator detectors were used for rejection of cosmic rays as well as charged particles from the target. BUNI was set at $`135^{}`$ and subtended a solid angle of 0.013 sr. Zero-degree calibrations of BUNI were done at the beginning and end of the experiment, in order to obtain both the lineshape of BUNI and an energy calibration for the NaI core. The NaI quadrants were calibrated periodically with a radioactive source (Th-C). Figure 1 depicts the BUNI energy spectrum (E$`_\gamma ^{}`$), corrected for photon absorption effects ($``$3%), from the free proton. Random coincidences as well as contributions from empty-target backgrounds (36% for $`(\gamma ,\gamma )`$ and 5% for $`(\gamma ,\gamma p)`$) have been subtracted. The spectra are summed over all tagged photon energies and shifted to the maximum incident energy of 260 MeV, with the appropriate kinematic corrections to account for the different scattered photon energy at each incident photon energy. Figure 1(a) is the energy spectrum from a free proton, without detection of the recoil proton. The Compton scattered events (the peak near 175 MeV) are clearly distinguishable from the large number of events from neutral pion ($`\pi ^0`$) production (below $``$165 MeV). The histogram is the result of a simulation of the experiment. The final yield was determined by integrating the data from 170 to 190 MeV and correcting for the Compton events excluded from this region (18%) and the $`\pi ^0`$ events included in this region (1%). For E$`{}_{\gamma }{}^{}=247`$ MeV and $`\theta _\gamma =135^{}`$, the cross section for elastic Compton scattering from the free proton ($`p(\gamma ,\gamma )p`$) was determined to be $$\frac{d\sigma }{d\mathrm{\Omega }_\gamma }=94.6\pm 9.0\pm 3.8\mathrm{nb}/\mathrm{sr},$$ (8) where the first error is statistical and the second is systematic. This is in good agreement with recent measurements . The dispersion calculations of L’vov et al. predict a somewhat lower value of 78.0 nb/sr using $`(\overline{\alpha }\overline{\beta })_p`$=10. The efficiency for the detection of protons in the array was determined in situ from the fraction of the $`p(\gamma ,\gamma )p`$ events (Fig. 1(a)) that included a detected proton (Fig. 1(b)). This correction ($`0.58\pm 0.02`$) was accounted for by a combination of the simulation and cuts in the analysis. For Compton scattering from a bound nucleon, the Fermi motion causes overlap in E$`_\gamma ^{}`$ between the Compton and $`\pi ^0`$ events. In order to obtain better separation, the missing mass ($`\mathrm{M}_\mathrm{X}`$) of the spectator nucleon was calculated. Figure 2 depicts the difference between the missing mass and the spectator mass ($`\mathrm{\Delta }\mathrm{M}_{\mathrm{miss}}`$) for the two quasi-free reactions. The yields are corrected for random coincidences, empty-target backgrounds (12–14%), and detection efficiencies. The histograms are from simulations of the experiment. The thin lines are the individual contributions from the $`\pi ^0`$ and Compton reactions while the thick line is the sum. The simulation assumes that the deuteron reactions proceed via quasi-free kinematics. The spectator nucleon is distributed according to its Fermi distribution and the angular distribution of the $`\pi ^0`$ or photon is determined by the theoretical calculations of Levchuk et al. The simulation gives a good fit to the data, and is relatively insensitive to the input angular distribution ($``$2%). The $`\chi ^2/N_{d.o.f.}`$ is 1.4 (proton) and 1.5 (neutron) for $`N_{d.o.f.}`$=63 ($`20`$ to 110 MeV). To extract the yields for the Compton reactions, the data were integrated from $`20`$ to +20 MeV and corrected for the Compton events excluded from this region (10–15%) and the $`\pi ^0`$ events included in this region (7–8%). In order to compare to theoretical predictions, the measured cross sections were interpolated to the quasi-free kinematic point ($`P_{spectator}=0`$) through use of the simulation. The average and root-mean-square values for $`P_{spectator}`$ were 48 and 55 MeV/c, respectively. For E$`{}_{\gamma }{}^{}=247`$ MeV and $`\theta _\gamma ^{}=135^{}`$, the final differential cross sections, at the quasi-free point, are $`{\displaystyle \frac{d^3\sigma }{d\mathrm{\Omega }_\gamma d\mathrm{\Omega }_NdE_N}}`$ $`=`$ $`56.5\pm 2.8\pm 4.5(\mathrm{proton}),`$ (9) $`=`$ $`33.3\pm 7.2\pm 3.0(\mathrm{neutron}),`$ (10) where the units are nb/sr<sup>2</sup>/MeV, the first error is statistical, and the second is systematic. The proton result is consistent with the recent results of Wissmann et al. . Since the back-angle Compton cross section is primarily sensitive to the difference of the polarizabilities, $`(\overline{\alpha }\overline{\beta })_p`$ was used as a fit parameter in matching the theoretical calculation Levchuk et al. to the bound proton data (Fig. 3(a)). The solid line is the theoretical calculation, the thin line is the central fitted value, and the horizontal dashed lines indicate the combined-error band (the quadratic sum of statistical and systematic). The model dependence of the calculation is small compared to the experimental errors. The fitted value, $$(\overline{\alpha }\overline{\beta })_p=14.7_{4.0}^{+4.6},$$ (11) is consistent with the free proton result (Eq. 1), as well as with the quasi-free results of Wissmann et al. . This indicates that the theoretical calculations are adequately describing the nuclear effects. The theoretical calculations have been carried out using a ‘traditional’ value for the backward-spin polarizability of $`\gamma _\pi `$=$`36.8`$. Using the value $`\gamma _\pi `$=$`27.1`$, suggested by Tonnison et al. , results in a fitted value of $$(\overline{\alpha }\overline{\beta })_p=26.6_{5.8}^{+7.8},$$ (12) which is over two standard deviations from the free value. Our data show no evidence for modification of $`\gamma _\pi `$, in agreement with the conclusions of Wissmann et al. . The ratio of the bound proton to the bound neutron quasi-free cross section is 1.70$`\pm `$0.38$`\pm `$0.13. Using the ratio of the measured cross sections minimizes the effect of systematic errors on the extraction of $`(\overline{\alpha }\overline{\beta })_n`$, and reduces the model dependence inherent in extracting information on the free-neutron from quasi-free results. The ratio for the theoretical calculation was obtained by fixing the proton polarizabilities (Eq. 1) and treating $`(\overline{\alpha }\overline{\beta })_n`$ as a fit parameter (Fig. 3(b)). The best fit was obtained with $`(\overline{\alpha }\overline{\beta })_n=12`$, with a lower limit of $`0`$. No upper limit can be obtained. The model uncertainty in extracting $`(\overline{\alpha }\overline{\beta })_n`$ is expected to be negligible compared to the experimental error. The central value is markedly different from the value extracted from elastic Compton scattering from the deuteron (Eq. 6). However, the two are within errors. Invoking the sum rule , the values for the polarizabilities are $`\overline{\alpha }_n=13.6`$, with a lower limit of 7.6, and $`\overline{\beta }_n`$ = 1.6, with an upper limit of 7.6. The upper (lower) limits on $`\overline{\alpha }_n`$ ($`\overline{\beta }_n`$) are unconstrained by the present measurement. Combining our results with those from Rose et al. , we obtain one-sigma constraints of $`\overline{\alpha }_n=7.614.0`$ and $`\overline{\beta }_n=1.27.6`$. These can be compared with recent Chiral Perturbation Theory (ChPT) predictions at $`O(Q^4)`$ $$\overline{\alpha }_n=13.4\pm 1.5,\overline{\beta }_n=7.8\pm 3.6.$$ (13) From Eqs. 7 and 9 the effective free-proton cross section at $`\theta _\gamma ^{}=135^{}`$ can be determined from that of the bound proton ($`d(\gamma ,\gamma ^{}p)n`$). The result is $$\frac{d\sigma (\gamma p\gamma ^{}p)}{d\mathrm{\Omega }_\gamma }=66.8\pm 3.4\pm 5.3\mathrm{nb}/\mathrm{sr},$$ (14) corresponding to E$`{}_{\gamma }{}^{}=244.6`$ MeV for the free proton. Adjusting Eq. 8 to this lower energy results in a $``$5% drop in the measured cross section to 90.1$`\pm `$8.6$`\pm `$3.6 nb/sr. The ratio of the measured to effective free values is $`1.35\pm 0.18`$, indicating they are not in good agreement. Model-dependence in extracting effective free-neutron values from quasi-free should be minimized by using the ratio of quasi-free cross sections, as discussed above. The dispersion calculations of L’vov et al. predict a free value of 74.2 nb/sr for $`(\overline{\alpha }\overline{\beta })_p`$=10. Eqs. 7 and 10 can also be used to determine the effective free cross section for the neutron at $`\theta _\gamma ^{}=135^{}`$: $$\frac{d\sigma (\gamma n\gamma ^{}n)}{d\mathrm{\Omega }_\gamma }=39.5\pm 8.5\pm 3.6\mathrm{nb}/\mathrm{sr},$$ (15) corresponding to E$`{}_{\gamma }{}^{}=244.6`$ MeV for the free neutron. This is the first experimental determination of the cross section for Compton scattering from the free neutron. An alternate method for extracting the free-neutron cross section from the measured values is to take the ratio of free to quasi-free cross sections for the proton (Eqs. 8 and 9) and multiply by the quasi-free neutron cross section (Eq. 10), resulting in 55.8$`\pm `$13.5$`\pm `$4.9. The ratio of the extracted free values $`1.41\pm 0.47`$ is consistent with unity, albeit with large errors. The dispersion calculations of L’vov et al. predict a free value of 45.7 nb/sr (for $`(\overline{\alpha }\overline{\beta })_n`$=10) in agreement with the experimental results. In summary, the polarizabilities of the bound proton and neutron have, for the first time, been simultaneously extracted from quasi-free Compton scattering measurements. The values for the neutron (Eq. 13) are consistent with those known for the free proton (Eq. 5), as expected from charge-symmetry arguments, and are in accord with recent ChPT predictions. The authors would like to thank M.I. Levchuk, A.I. L’vov, and V.A. Petrun’kin for supplying the results from their theoretical calculations. N.R.K. would like to thank M.I.L. for many useful discussions. This work was supported in part by grants from the Natural Science and Engineering Research Council of Canada and the National Science Foundation (USA).
warning/0003/math0003120.html
ar5iv
text
# The automorphism tower problem revisited ## 1. Introduction If $`G`$ is a centreless group, then there is a natural embedding of $`G`$ into its automorphism group $`\mathrm{Aut}G`$, obtained by sending each $`gG`$ to the corresponding inner automorphism $`i_g\mathrm{Aut}G`$. In this paper, we will always work with the left action of $`\mathrm{Aut}G`$ on $`G`$. Thus $`i_g(x)=gxg^1`$ for all $`xG`$. If $`\pi \mathrm{Aut}G`$ and $`gG`$, then an easy calculation shows that $`\pi i_g\pi ^1=i_{\pi (g)}`$. Hence the group of inner automorphisms $`\mathrm{Inn}G`$ is a normal subgroup of $`\mathrm{Aut}G`$; and $`C_{\mathrm{Aut}G}(\mathrm{Inn}G)=1`$. In particular, $`\mathrm{Aut}G`$ is also a centreless group. This enables us to define the automorphism tower of $`G`$ to be the ascending chain of groups $$G=G_0\mathrm{}G_1\mathrm{}G_2\mathrm{}\mathrm{}G_\alpha \mathrm{}G_{\alpha +1}\mathrm{}\mathrm{}$$ such that for each ordinal $`\alpha `$ 1. $`G_{\alpha +1}=\mathrm{Aut}G_\alpha `$; and 2. if $`\alpha `$ is a limit ordinal, then $`G_\alpha =\underset{\beta <\alpha }{}G_\beta `$. (At each successor step, we identify $`G_\alpha `$ with $`\mathrm{Inn}G_\alpha `$ via the natural embedding.) The automorphism tower is said to terminate if there exists an ordinal $`\alpha `$ such that $`G_{\alpha +1}=G_\alpha `$. This occurs if and only if there exists an ordinal $`\alpha `$ such that $`\mathrm{Aut}G_\alpha =\mathrm{Inn}G_\alpha `$. A classical result of Wielandt says that if $`G`$ is finite, then the automorphism tower terminates after finitely many steps. Wielandt’s theorem fails for infinite centreless groups. For example, consider the infinite dihedral group $`D_{\mathrm{}}=a,b`$, where $`a`$ and $`b`$ are elements of order 2. Then $`D_{\mathrm{}}=ab`$ is the free product of its cyclic subgroups $`a`$ and $`b`$. It follows that $`D_{\mathrm{}}`$ is a centreless group; and that $`D_{\mathrm{}}`$ has an outer automorphism $`\pi `$ of order 2 which interchanges the elements $`a`$ and $`b`$. It is easily shown that $`\mathrm{Aut}D_{\mathrm{}}=\pi ,i_a`$. Thus $`\mathrm{Aut}D_{\mathrm{}}`$ is also an infinite dihedral group, and so $`\mathrm{Aut}D_{\mathrm{}}D_{\mathrm{}}`$. Hence for each $`n\omega `$, the $`n^{th}`$ group in the automorphism tower of $`D_{\mathrm{}}`$ is isomorphic to $`D_{\mathrm{}}`$; and the automorphism tower of $`D_{\mathrm{}}`$ does not terminate after finitely many steps. In the 1970s, a number of special cases of the automorphism tower problem were solved. For example, Rae and Roseblade proved that the automorphism tower of a centreless Černikov group terminates after finitely many steps; and Hulse proved that the automorphism tower of a centreless polycyclic group terminates in countably many steps. But the problem was not solved in full generality until 1984, when Thomas showed that the automorphism tower of an arbitrary centreless group eventually terminates; and that for each ordinal $`\alpha `$, there exists a group whose automorphism tower terminates in exactly $`\alpha `$ steps. ###### Definition 1.1. If $`G`$ is a centreless group, then the height $`\tau (G)`$ of the automorphism tower of $`G`$ is the least ordinal $`\alpha `$ such that $`G_{\alpha +1}=G_\alpha `$. This raises the question of finding bounds for $`\tau (G)`$ in terms of the cardinality of $`G`$. In his original paper , Thomas proved that if $`G`$ is an infinite centreless group of cardinality $`\kappa `$, then $`\tau (G)\left(2^\kappa \right)^+`$. Soon afterwards, Thomas and Felgner independently noticed that an easy application of Fodor’s Lemma yielded the following slightly better bound. ###### Theorem 1.2 (Thomas ). If $`G`$ is an infinite centreless group of cardinality $`\kappa `$, then $`\tau (G)<\left(2^\kappa \right)^+`$. $`\mathrm{}`$ ###### Definition 1.3. If $`\kappa `$ is an infinite cardinal, then $`\tau _\kappa `$ is the least ordinal such that $`\tau (G)<\tau _\kappa `$ for every centreless group $`G`$ of cardinality $`\kappa `$. Since there are only $`2^\kappa `$ centreless groups of cardinality $`\kappa `$ up to isomorphism, it follows that $`\tau _\kappa <\left(2^\kappa \right)^+`$. On the other hand, Thomas has shown that for each ordinal $`\alpha <\kappa ^+`$, there exists a centreless group $`G`$ of cardinality $`\kappa `$ such that $`\tau (G)=\alpha `$. Thus $`\kappa ^+\tau _\kappa <\left(2^\kappa \right)^+`$. It is natural to ask whether a better explicit bound on $`\tau _\kappa `$ can be proved in $`ZFC`$, preferably one which does not involve cardinal exponentiation. The proof of Theorem 1.2 is extremely simple, and uses only the most basic results in group theory, together with some elementary properties of the infinite cardinal numbers. So it is not surprising that Theorem 1.2 does not give the best possible bound for $`\tau _\kappa `$. In contrast, the proof of Wielandt’s theorem is much deeper, and involves an intricate study of the subnormal subgroups of a finite centreless group. The real question behind the search for better explicit bounds for $`\tau _\kappa `$ is whether there exists a subtler, more informative, group-theoretic proof of the automorphism tower theorem for infinite groups. The main result of this paper says that no such bounds can be proved in $`ZFC`$, and thus can be interpreted as saying that no such proof exists. (It is perhaps worth mentioning that the proof of Theorem 1.2 yields that the automorphism tower of a finite centreless group terminates in countably many steps. However, there does not seem to be an easy reduction from countable to finite; and it appears that some form of Wielandt’s analysis is necessary.) ###### Theorem 1.4. Let $`VGCH`$ and let $`\kappa `$, $`\lambda V`$ be uncountable cardinals such that $`\kappa <\mathrm{cf}(\lambda )`$. Let $`\alpha `$ be any ordinal such that $`\alpha <\lambda ^+`$. Then there exists a notion of forcing $``$, which preserves cofinalities and cardinalities, such that the following statements are true in the corresponding generic extension $`V^{}`$. 1. $`2^\kappa =\lambda `$. 2. There exists a centreless group $`G`$ of cardinality $`\kappa `$ such that $`\tau (G)=\alpha `$. Thus it is impossible to find better explicit bounds for $`\tau _\kappa `$ when $`\kappa `$ is an uncountable cardinal. However, our methods do not enable us to deal with countable groups; and it remains an open question whether or not there exists a countable centreless group $`G`$ such that $`\tau (G)\omega _1`$. Most of this paper will be concerned with the problem of constructing centreless groups with extremely long automorphism towers. Unfortunately it is usually very difficult to compute the successive groups in an automorphism tower. We will get around this difficulty by reducing it to the much easier computation of the successive normalisers of a subgroup $`H`$ of a group $`G`$. ###### Definition 1.5. If $`H`$ is a subgroup of the group $`G`$, then the normaliser tower of $`H`$ in $`G`$ is defined inductively as follows. 1. $`N_0(H)=H`$. 2. If $`\alpha =\beta +1`$, then $`N_\alpha (H)=N_G\left(N_\beta (H)\right)`$. 3. If $`\alpha `$ is a limit ordinal, then $`N_\alpha (H)=\underset{\beta <\alpha }{}N_\beta (H)`$. It is sometimes necessary for the notation to include an explicit reference to the ambient group $`G`$. In this case, we will write $`N_\alpha (H)=N_\alpha (H,G)`$. As we will see in Section 2, if $`\alpha `$ is any ordinal, then it is easy to construct examples of pairs of groups, $`H<G`$, such that the normaliser tower of $`H`$ in $`G`$ terminates in exactly $`\alpha `$ steps. The following lemma, which was essentially proved in , will enable us to convert normaliser towers into corresponding automorphism towers. ###### Lemma 1.6. Let $`K`$ be a field such that $`|K|>3`$ and let $`H`$ be a subgroup of $`\mathrm{Aut}K`$. Let $$G=PGL(2,K)HP\mathrm{\Gamma }L(2,K)=PGL(2,K)\mathrm{Aut}K.$$ Then $`G`$ is a centreless group; and for each $`\alpha `$, $`G_\alpha =PGL(2,K)N_\alpha (H)`$, where $`N_\alpha (H)`$ is the $`\alpha ^{th}`$ group in the normaliser tower of $`H`$ in $`\mathrm{Aut}K`$. $`\mathrm{}`$ It is well-known that every group $`G`$ can be realised as the automorphism group of a suitable graph $`\mathrm{\Gamma }`$. Thus the following result implies that every group $`G`$ can also be realised as the automorphism group of a suitable field $`K`$. ###### Lemma 1.7 (Fried and Kollár ). Let $`\mathrm{\Gamma }=X,E`$ be any graph. Then there exists a field $`K_\mathrm{\Gamma }`$ of cardinality $`\mathrm{max}\{\left|X\right|,\omega \}`$ which satisfies the following conditions. 1. $`X`$ is an $`\mathrm{Aut}K_\mathrm{\Gamma }`$-invariant subset of $`K_\mathrm{\Gamma }`$. 2. The restriction mapping, $`\pi \pi X`$, is an isomorphism from $`\mathrm{Aut}K_\mathrm{\Gamma }`$ onto $`\mathrm{Aut}\mathrm{\Gamma }`$. $`\mathrm{}`$ Combining Lemmas 1.6 and 1.7, we obtain the following reduction of our problem. ###### Lemma 1.8. Suppose that there exists a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ and a subgroup $`H`$ of $`\mathrm{Aut}\mathrm{\Gamma }`$ such that 1. $`\left|H\right|\kappa `$; and 2. the normaliser tower of $`H`$ in $`\mathrm{Aut}\mathrm{\Gamma }`$ terminates in exactly $`\alpha `$ steps. Then there exists a centreless group $`T`$ of cardinality $`\kappa `$ such that $`\tau (T)=\alpha `$. ###### Proof. Let $`K_\mathrm{\Gamma }`$ be the corresponding field, which is given by Lemma 1.7, and let $`T=PGL(2,K_\mathrm{\Gamma })H`$. By Lemma 1.6, $`\tau (T)=\alpha `$. ∎ From now on, fix a regular uncountable cardinal $`\kappa `$ such that $`\kappa ^{<\kappa }=\kappa `$ and an ordinal $`\alpha `$. Roughly speaking, our strategy will be to 1. first construct a pair of groups, $`H<G`$, such that $`\left|H\right|\kappa `$ and the normaliser tower of $`H`$ in $`G`$ terminates in $`\alpha `$ steps; and 2. then attempt to find a cardinal-preserving notion of forcing $``$ which adjoins a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`G\mathrm{Aut}\mathrm{\Gamma }`$. Of course, there are many groups $`G`$ for which such a notion of forcing $``$ cannot possibly exist. For example, De Bruijn has shown that the alternating group $`\mathrm{Alt}(\kappa ^+)`$ cannot be embedded in $`\mathrm{Sym}(\kappa )`$. Consequently, there is no cardinal-preserving notion of forcing $``$ which adjoins a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`\mathrm{Alt}(\kappa ^+)\mathrm{Aut}\mathrm{\Gamma }`$. Our next definition singles out a combinatorial condition which is satisfied by all those groups $`G`$ such that $`G`$ is embeddable in $`\mathrm{Sym}(\kappa )`$. (See Proposition 1.11.) Conversely, in Theorem 1.12, we will show that if a group $`G`$ satisfies this combinatorial condition, then there exists a cardinal-preserving notion of forcing $``$ which adjoins a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`G\mathrm{Aut}\mathrm{\Gamma }`$. ###### Definition 1.9. Let $`\kappa `$ be a regular uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$. Then a group $`G`$ is said to satisfy the $`\kappa ^+`$-compatibility condition if it has the following property. Suppose that $`H`$ is a group such that $`\left|H\right|<\kappa `$. Suppose that $`\left(f_ii<\kappa ^+\right)`$ is a sequence of embeddings $`f_i:HG`$; and let $`H_i=f_i[H]`$ for each $`i<\kappa ^+`$. Then there exist ordinals $`i<j<\kappa ^+`$ and a surjective homomorphism $`\phi :H_i,H_jH_i`$ such that 1. $`\phi f_j=f_i`$; and 2. $`\phi H_i=id_{H_i}`$. ###### Example 1.10. To get an understanding of Definition 1.9, it will probably be helpful to see an example of a group which fails to satisfy the $`\kappa ^+`$-compatibility condition. So we will show that $`\mathrm{Alt}(\kappa ^+)`$ does not satisfy the $`\kappa ^+`$-compatibility condition. Let $`H=\mathrm{Alt}(4)`$. For each $`3i<\kappa ^+`$, let $`\mathrm{\Delta }_i=\{0,1,2,i\}`$ and let $`f_i:\mathrm{Alt}(4)\mathrm{Alt}(\mathrm{\Delta }_i)`$ be an isomorphism. If $`3i<j<\kappa ^+`$, then $$\mathrm{Alt}(\mathrm{\Delta }_i),\mathrm{Alt}(\mathrm{\Delta }_j)=\mathrm{Alt}(\mathrm{\Delta }_i\mathrm{\Delta }_j)\mathrm{Alt}(5).$$ Since $`\mathrm{Alt}(5)`$ is a simple group, there does not exist a surjective homomorphism from $`\mathrm{Alt}(\mathrm{\Delta }_i),\mathrm{Alt}(\mathrm{\Delta }_j)`$ onto $`\mathrm{Alt}(\mathrm{\Delta }_i)`$. ###### Proposition 1.11. Let $`\kappa `$ be a regular uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$, and let $`G\mathrm{Sym}(\kappa )`$. Then $`G`$ satisfies the $`\kappa ^+`$-compatibility condition. ###### Proof. Let $`H`$ be a group such that $`\left|H\right|<\kappa `$, and let $`\left(f_ii<\kappa ^+\right)`$ be a sequence of embeddings $`f_i:HG`$. For each $`i<\kappa ^+`$, let $`H_i=f_i[H]`$; and let $`Z_i`$ be a subset of $`\kappa `$ chosen so that 1. $`\left|Z_i\right|<\kappa `$; 2. $`Z_i`$ is $`H_i`$-invariant; and 3. $`gZ_iid_{Z_i}`$ for all $`1gH_i`$. After passing to a suitable subsequence if necessary, we can suppose that the following conditions hold. 1. There exists a fixed subset $`Z`$ such that $`Z_i=Z`$ for all $`i<\kappa ^+`$. 2. For each $`i<\kappa ^+`$, let $`r_i:H_i\mathrm{Sym}(Z)`$ be the restriction mapping, $`ggZ`$; and let $`\pi _i:H\mathrm{Sym}(Z)`$ be the embedding defined by $`\pi _i=r_if_i`$. Then $`\pi _i=\pi _j`$ for all $`i<j<\kappa ^+`$. Fix any pair of ordinals $`i`$, $`j`$ such that $`i<j<\kappa ^+`$. Let $`\rho :H_i,H_j\mathrm{Sym}(Z)`$ be the restriction mapping, $`ggZ`$. Then $`\rho \left[H_i,H_j\right]=r_i\left[H_i\right]`$, and so we can define a surjective homomorphism $`\phi :H_i,H_jH_i`$ by $`\phi =r_i^1\rho `$. Clearly $`\phi H_i=id_{H_i}`$; and it is easily checked that $`\phi f_j=f_i`$. ∎ ###### Theorem 1.12. Let $`\kappa `$ be a regular uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$, and let $`G`$ be a group which satisfies the $`\kappa ^+`$-compatibility condition. Then there exists a notion of forcing $``$ such that 1. $``$ is $`\kappa `$-closed; 2. $``$ has the $`\kappa ^+`$-$`c.c.`$; and 3. $`\underset{}{}\text{ There exists a graph }\mathrm{\Gamma }\text{ of cardinality }\kappa \text{ such that }G\mathrm{Aut}\mathrm{\Gamma }`$. Furthermore, if $`\left|G\right|=\lambda `$, then $`\left|\right|=\mathrm{max}\{\kappa ,\lambda ^{<\kappa }\}`$. Combining Proposition 1.11 and Theorem 1.12, we see that if $`\kappa `$ is an uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$ and $`G`$ is an arbitrary subgroup of $`\mathrm{Sym}(\kappa )`$, then there exists a cardinal-preserving notion of forcing $``$ and a graph $`\mathrm{\Gamma }V^{}`$ such that $`G\mathrm{Aut}\mathrm{\Gamma }`$. This result is not true of arbitrary subgroups of $`\mathrm{Sym}(\omega )`$; for Solecki has shown that no uncountable free abelian group is the automorphism group of a countable first-order structure. Theorem 1.12 will be proved in Section 3. It is now easy to explain the main points of the proof of Theorem 1.4. Assume that $`VGCH`$. Let $`\kappa `$ be a regular uncountable cardinal, and let $`\lambda `$ be a cardinal such that $`\mathrm{cf}(\lambda )>\kappa `$. Let $`\alpha `$ be any ordinal such that $`\alpha <\lambda ^+`$. In Section 2, we will prove that there exists a notion of forcing $``$ such that 1. $``$ is $`\kappa `$-closed; 2. $``$ has the $`\kappa ^+`$-$`c.c.`$; and such that the following statements are true in the generic extension $`V^{}`$. 1. $`2^\kappa =\lambda `$. 2. There exist groups $`H<G<\mathrm{Sym}(\kappa )`$ such that $`|H|=\kappa `$ and the normaliser tower of $`H`$ in $`G`$ terminates in exactly $`\alpha `$ steps. By Proposition 1.11, $`G`$ satisfies the $`\kappa ^+`$-compatibility condition. Hence we can use Theorem 1.12 to generically adjoin a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`G\mathrm{Aut}\mathrm{\Gamma }`$. A moment’s thought shows that the normaliser tower of $`H`$ in $`G`$ is an absolute notion. Thus Lemma 1.8 yields a centreless group $`T`$ of cardinality $`\kappa `$ such that $`\tau (T)=\alpha `$. The case when $`\kappa `$ is a singular cardinal requires a little more work, and will be dealt with in Section 4. The remainder of this section will be devoted to another two easy applications of Theorem 1.12. ###### Application 1.13. A well-known open problem asks whether there exists a countable structure $``$ such that $`\mathrm{Aut}`$ is the free group on $`2^\omega `$ generators. Using Theorem 1.12, it is easy to establish the consistency of the existence of a structure $`𝒩`$ of cardinality $`\omega _1`$ such that $`\mathrm{Aut}𝒩`$ is the free group on $`2^{\omega _1}`$ generators. It is not known whether the existence of such a structure can be proved in $`ZFC`$. ###### Theorem 1.14. Let $`V`$ be a transitive model of $`ZFC`$ and let $`\kappa `$, $`\lambda `$, $`\theta `$ be cardinals such that $`\kappa ^{<\kappa }=\kappa <\lambda \theta =\theta ^\kappa `$. Then there exists a notion of forcing $``$, which preserves cofinalities and cardinalities, such that the following statements are true in $`V^{}`$. 1. $`2^\kappa =\theta `$; and 2. there exists a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`\mathrm{Aut}\mathrm{\Gamma }`$ is the free group on $`\lambda `$ generators. ###### Proof. After performing a preliminary forcing if necessary, we can also suppose that $`2^\kappa =\theta `$. Let $`F`$ be the free group on $`\lambda `$ generators. Then it is enough to show that $`F`$ satisfies the $`\kappa ^+`$-compatibility condition. Let $`H`$ be a (necessarily free) group such that $`\left|H\right|<\kappa `$, and let $`\left(f_ii<\kappa ^+\right)`$ be a sequence of embeddings $`f_i:HF`$. For each $`i<\kappa ^+`$, let $`H_i=f_i[H]`$. Then $`H_ii<\kappa ^+`$ is a free group of cardinality at most $`\kappa ^+`$. By Theorem 5.1 , there exists an embedding of $`H_ii<\kappa ^+`$ into $`\mathrm{Sym}(\kappa )`$. So Proposition 1.11 yields the existence of ordinals $`i<j<\kappa ^+`$ and a surjective homomorphism $`\phi :H_i,H_jH_i`$ such that $`\phi f_j=f_i`$ and $`\phi H_i=id_{H_i}`$. ∎ ###### Application 1.15. Theorem 1.6 says that if $`G`$ is a finitely generated centreless group, then the automorphism tower of $`G`$ terminates in countably many steps. It is conceivable that a more general result holds; namely, that the automorphism tower of $`G`$ terminates in countably many steps, whenever $`G`$ is a countable centreless group such that $`\mathrm{Aut}G`$ is also countable. To see why this might be true, let $`G`$ be such a group. Then, by Kueker , there exists a finite subset $`FG`$ such that each automorphism $`\pi \mathrm{Aut}G`$ is uniquely determined by its restriction $`\pi F`$. In terms of the automorphism tower of $`G`$, this says that there is a finite subset $`FG`$ such that $`C_{G_1}(F)=1`$. Suppose that the “rigidity” of $`F`$ within $`G=G_0`$ is propagated along the automorphism tower of $`G`$; ie. that $`C_{G_\alpha }(F)=1`$ for all ordinals $`\alpha <\omega _1`$. Then the proof of Theorem 1.6 shows that the automorphism tower of $`G`$ terminates in countably many steps. ###### Question 1.16. Let $`G`$ be a centreless group such that $`\left|\mathrm{Aut}G\right|=\omega `$. Does there exist a finite subset $`FG`$ such that $`C_{G_\alpha }(F)=1`$ for all ordinals $`\alpha <\omega _1`$? If $`\left|G_\alpha \right|=\omega `$ for all $`\alpha <\omega _1`$, then Fodor’s Lemma implies that there exists an ordinal $`\beta <\omega _1`$ and a finite subset $`F_\beta `$ of $`G_\beta `$ such that $`C_{G_\alpha }(F_\beta )=1`$ for all $`\beta \alpha <\omega _1`$; and so $`\tau (G)<\omega _1`$. This observation suggests the following weak form of Question 1.16, which is also open. ###### Question 1.17. Does there exist a centreless group $`G`$ such that $`\left|\mathrm{Aut}G\right|=\omega `$ and $`\left|\mathrm{Aut}(\mathrm{Aut}G)\right|=2^\omega `$? Of course, a positive answer to Question 1.16 implies a negative answer to Question 1.17. Using Theorem 1.12, it is easy to establish the consistency of the existence of a centreless group $`G`$ of cardinality $`\omega _1`$ such that $`\left|\mathrm{Aut}G\right|=\omega _1`$ and $`\left|\mathrm{Aut}(\mathrm{Aut}G)\right|=2^{\omega _1}`$. Once again, it is not known whether the existence of such a group can be proved in $`ZFC`$. ###### Theorem 1.18. Let $`\kappa `$ be a regular uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$. Then it is consistent that there exists a centreless group $`G`$ of cardinality $`\kappa `$ such that $`\left|\mathrm{Aut}G\right|=\kappa `$ and $`\left|\mathrm{Aut}(\mathrm{Aut}G)\right|=2^\kappa `$. ###### Proof. Let $`V`$ be the ground model. For each $`\alpha `$, $`\xi <\kappa `$, let $`Z_\xi ^\alpha =z_\xi ^\alpha `$ be an infinite cyclic group. For each $`\xi <\kappa `$, let $`A_\xi =\underset{\alpha <\kappa }{}Z_\xi ^\alpha `$; and let $`B=\underset{\xi <\kappa }{}A_\xi `$. Define an action of $`\mathrm{Sym}(\kappa )`$ on $`B`$ by $`\pi z_\xi ^\alpha \pi ^1=z_{\pi (\xi )}^\alpha `$ for all $`\alpha `$, $`\xi <\kappa `$; and let $`W=B\mathrm{Sym}(\kappa )`$ be the corresponding semidirect product. Let $`H=\underset{\xi <\kappa }{}Z_\xi ^\xi `$. Then the members of the normaliser tower of $`H`$ in $`W`$ are 1. $`N_0(H)=H`$; 2. $`N_1(H)=B`$; 3. $`N_2(H)=W`$. Clearly $`W`$ is embeddable in $`\mathrm{Sym}(\kappa )`$; and so $`W`$ satisfies the $`\kappa ^+`$-compatibility condition. Let $``$ be the notion of forcing, given by Theorem 1.12, which adjoins a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`W\mathrm{Aut}\mathrm{\Gamma }`$. Let $`K_\mathrm{\Gamma }V^{}`$ be the corresponding field, which is given by Lemma 1.7. Then $`G=PGL(2,K_\mathrm{\Gamma })H`$ is a group such that $$\left|\mathrm{Aut}G\right|=\left|PGL(2,K_\mathrm{\Gamma })B\right|=\kappa $$ and $$\left|\mathrm{Aut}(\mathrm{Aut}G)\right|=\left|PGL(2,K_\mathrm{\Gamma })W\right|=2^\kappa .$$ Our set-theoretic notation mainly follows that of Kunen . Thus if $``$ is a notion of forcing and $`p`$,$`q`$, then $`qp`$ means that $`q`$ is a strengthening of $`p`$. We say that $``$ is $`\kappa `$-closed if for every $`\lambda <\kappa `$, every descending sequence of elements of $``$ $$p_0p_1\mathrm{}p_\xi \mathrm{},\xi <\lambda ,$$ has a lower bound in $``$. If $`V`$ is the ground model, then we will denote the generic extension by $`V^{}`$ if we do not wish to specify a particular generic filter $`H`$. If we want to emphasize that the term $`t`$ is to be interpreted in the model $`M`$ of $`ZFC`$, then we write $`t^M`$; for example, $`\mathrm{Sym}(\lambda )^M`$. Our group-theoretic notation is standard. For example, the (restricted) wreath product of $`A`$ by $`C`$ is denoted by $`A\text{wr}C`$; and the direct sum of the groups $`H_\xi `$, $`\xi <\lambda `$, is denoted by $`\underset{\xi <\lambda }{}H_\xi `$. If $`\pi \mathrm{Sym}(\kappa )`$, then $`\mathrm{supp}(\pi )=\{\alpha \kappa \pi (\alpha )\alpha \}`$; and if $`\lambda `$ is an infinite cardinal such that $`\lambda \kappa `$, then $`\mathrm{Sym}_\lambda (\kappa )=\{\pi \mathrm{Sym}(\kappa )\left|\mathrm{supp}(\pi )\right|<\lambda \}`$. ## 2. Realising normaliser towers within infinite symmetric groups Let $`\kappa `$ be a regular uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$. In this section, we will study the problem of realising long normaliser towers within $`\mathrm{Sym}(\kappa )`$. In particular, we will prove that if $`\alpha `$ is any ordinal, then there exists a generic extension $`V^{}`$ such that a normaliser tower of height $`\alpha `$ can be realised in $`\mathrm{Sym}(\kappa )^V^{}`$. First for each ordinal $`\alpha `$, we will construct a pair of groups, $`H<G`$, such that the normaliser tower of $`H`$ in $`G`$ terminates in exactly $`\alpha `$ steps. ###### Definition 2.1. The ascending chain of groups $$W_0W_1\mathrm{}W_\alpha W_{\alpha +1}\mathrm{}$$ is defined inductively as follows. 1. $`W_0=C_2`$, the cyclic group of order 2. 2. Suppose that $`\alpha =\beta +1`$. Then $$W_\beta =W_\beta 1\left[W_\beta W_\beta ^{}\right]\sigma _{\beta +1}=W_{\beta +1}.$$ Here $`W_\beta ^{}`$ is an isomorphic copy of $`W_\beta `$; and $`\sigma _{\beta +1}`$ is an element of order 2 which interchanges the factors $`W_\beta 1`$ and $`1W_\beta ^{}`$ of the direct sum $`W_\beta W_\beta ^{}`$ via conjugation. Thus $`W_{\beta +1}`$ is isomorphic to the wreath product $`W_\beta \text{wr}C_2`$. 3. If $`\alpha `$ is a limit ordinal, then $`W_\alpha =\underset{\beta <\alpha }{}W_\beta `$. ###### Lemma 2.2. $`\left|W_\alpha \right|\mathrm{max}\{|\alpha |,\omega \}`$ for all ordinals $`\alpha `$. ###### Proof. This follows by an easy induction on $`\alpha `$. ∎ ###### Lemma 2.3. 1. If $`1n<\omega `$, then the normaliser tower of $`W_0`$ in $`W_n`$ terminates in exactly $`n+1`$ steps. 2. If $`\alpha \omega `$, then the normaliser tower of $`W_0`$ in $`W_\alpha `$ terminates in exactly $`\alpha `$ steps. ###### Proof. (a) It is easily checked that $$N_1(W_0,W_n)=W_0W_0^{}W_1^{}\mathrm{}W_{n1}^{}$$ and that $$N_2(W_0,W_n)=W_1W_1^{}\mathrm{}W_{n1}^{};$$ and that, in general, for each $`0\mathrm{}n1`$, $$N_{\mathrm{}+1}(W_0,W_n)=W_{\mathrm{}}\underset{\mathrm{}m<n}{}W_m^{}.$$ (b) For example, consider the case when $`\alpha >\omega `$. Then for each $`\mathrm{}\omega `$, $$N_{\mathrm{}+1}(W_0,W_\alpha )=W_{\mathrm{}}\underset{\mathrm{}\beta <\alpha }{}W_\beta ^{};$$ and for each $`\gamma `$ such that $`\omega \gamma <\alpha `$, $$N_\gamma (W_0,W_\alpha )=W_\gamma \underset{\gamma \beta <\alpha }{}W_\beta ^{}.$$ ###### Remark 2.4. Unfortunately, the group $`W_{\kappa ^+}`$ does not satisfy the $`\kappa ^+`$-compatibility condition. To see this, let $$H=C_2\text{wr}C_2=\left[ab\right]c;$$ and for each limit ordinal $`i<\kappa ^+`$, let $`f_i:HW_{\kappa ^+}`$ be the embedding such that $`f_i(a)=\sigma _{i+1}`$ and $`f_i(c)=\sigma _{i+2}`$. (Here we are using the notation which was introduced in Definition 2.1.) Let $`i`$, $`j`$ be any limit ordinals such that $`i<j<\kappa ^+`$. Then $$H_i,H_j\left(\left(C_2\text{wr}C_2\right)\text{wr}C_2\right)\text{wr}C_2.$$ Suppose that there exists a a surjective homomorphism $`\phi :H_i,H_jH_i`$ such that 1. $`\phi f_j=f_i`$; and 2. $`\phi H_i=id_{H_i}`$. Then $`\phi (\sigma _{j+1})=\phi (\sigma _{i+1})=\sigma _{i+1}`$ and $`\phi (\sigma _{j+2})=\phi (\sigma _{i+2})=\sigma _{i+2}`$. Consider the element $`x=\sigma _{j+1}\sigma _{j+2}\sigma _{j+1}\sigma _{j+2}H_j`$. Then it is easily checked that 1. $`x`$ lies in the centre of $`H_j`$; and 2. $`\sigma _{j+1}y\sigma _{j+1}^1=xyx^1`$ for all $`yH_i`$. Thus $`z=\phi (x)`$ lies in the centre of $`H_i`$. Since $`\phi (\sigma _{j+1})=\sigma _{i+1}`$, we find that $$\sigma _{i+1}y\sigma _{i+1}^1=zyz^1=y$$ for all $`yH_i`$. But this contradicts the fact that $`\sigma _{i+1}`$ is a noncentral element of $`H_i`$. Thus if $`\alpha \kappa ^+`$, then $`W_\alpha `$ is not embeddable in $`\mathrm{Sym}(\kappa )`$. However, the above argument does not rule out the possibility that $`W_\alpha `$ is embeddable in the quotient group $`\mathrm{Sym}(\kappa )/\mathrm{Sym}_\kappa (\kappa )`$; and this is enough for our purposes. ###### Lemma 2.5. Let $`\kappa `$ be an infinite cardinal such that $`\kappa ^{<\kappa }=\kappa `$. Suppose that $`\gamma `$ is an ordinal, and that there exists an embedding $$f:W_\gamma \mathrm{Sym}(\kappa )/\mathrm{Sym}_\kappa (\kappa ).$$ Then for each ordinal $`\alpha \gamma `$, there exist groups $`HG\mathrm{Sym}(\kappa )`$ such that $`|H|=\kappa `$ and the normaliser tower of $`H`$ in $`G`$ terminates in exactly $`\alpha `$ steps. ###### Proof. For each $`\alpha \gamma `$, let $`H^\alpha `$ be the subgroup of $`\mathrm{Sym}(\kappa )`$ such that $`f\left[W_\alpha \right]=H^\alpha /\mathrm{Sym}_\kappa (\kappa )`$. Since $`|\mathrm{Sym}_\kappa (\kappa )|=\kappa ^{<\kappa }=\kappa `$, it follows that $`|H^0|=\kappa `$. ###### Claim 2.6. For each ordinal $`\beta `$, $$f\left[N_\beta (W_0,W_\alpha )\right]=N_\beta (H^0,H^\alpha )/\mathrm{Sym}_\kappa (\kappa ).$$ ###### Proof. We will argue by induction on $`\beta `$. The result is clear when $`\beta =0`$, and no difficulties arise when $`\beta `$ is a limit ordinal. Suppose that $`\beta =\xi +1`$ and that the result holds for $`\xi `$. Let $`R=N_\xi (H^0,H^\alpha )`$; and for each subgroup $`K`$ such that $`\mathrm{Sym}_\kappa (\kappa )KH^\alpha `$, let $`\overline{K}=K/\mathrm{Sym}_\kappa (\kappa )`$. Then $$f\left[N_{\xi +1}(W_0,W_\alpha )\right]=N_{\overline{H^\alpha }}(\overline{R});$$ and so we must show that $$N_{\overline{H^\alpha }}(\overline{R})=\overline{N_{H^\alpha }(R)}.$$ But this is an immediate consequence of the Correspondence Theorem for subgroups of quotient groups, together with the observation that the normaliser of any subgroup $`L`$ is the largest subgroup $`M`$ such that $`L\mathrm{}M`$. ∎ It is now easy to complete the proof of Lemma 2.5. Applying Lemma 2.3 and Claim 2.6, we see that if $`\alpha \omega `$, then the normaliser tower of $`H^0`$ in $`H^\alpha `$ terminates in exactly $`\alpha `$ steps; and that if $`2\alpha =n<\omega `$, then the normaliser tower of $`H^0`$ in $`H^{n1}`$ terminates in exactly $`n`$ steps. This just leaves the cases when $`\alpha =0,1`$. When $`\alpha =0`$, we can take $`H=G=\mathrm{Alt}(\kappa )`$; and when $`\alpha =1`$, we can take $`H=\mathrm{Alt}(\kappa )`$ and $`G=\mathrm{Sym}(\kappa )`$. ∎ The next result implies that if $`\omega <\kappa ^\kappa =\kappa `$ and $`W`$ is any group, then there exists a cardinal-preserving notion of forcing $``$ such that in $`V^{}`$, the group $`W`$ is embeddable in $`\mathrm{Sym}(\kappa )/\mathrm{Sym}_\kappa (\kappa )`$. (If $`|W|^\kappa >|W|`$, then just embed $`W`$ in a larger group $`L`$ such that $`|L|^\kappa =|L|`$; and then apply Lemma 2.7 to $`L`$.) ###### Lemma 2.7. Let $`V`$ be a transitive model of $`ZFC`$ and let $`\kappa `$, $`\lambda `$ be cardinals such that $`\omega <\kappa ^{<\kappa }=\kappa <\lambda =\lambda ^\kappa `$. Let $`W`$ be any group of cardinality $`\lambda `$. Then there exists a notion of forcing $``$ such that 1. $``$ is $`\kappa `$-closed. 2. $``$ has the $`\kappa ^+`$-$`c.c.`$; and such that the following statements are true in the generic extension $`V^{}`$. 1. $`2^\kappa =\lambda `$. 2. There exists an isomorphic embedding $$f:W\left(\mathrm{Sym}(\kappa )/\mathrm{Sym}_\kappa (\kappa )\right)^V^{}.$$ ###### Proof. Let $`\mathrm{\Omega }=_{\alpha <\kappa }\{\alpha \}\times \alpha `$. We will work with the symmetric group $`\mathrm{Sym}(\mathrm{\Omega })`$ rather than with $`\mathrm{Sym}(\kappa )`$. Let $``$ be the notion of forcing consisting of the conditions $$p=(\delta _p,H_p,E_p)$$ such that the following hold. 1. $`\omega \delta _p<\kappa `$. 2. $`H_p`$ is a subgroup of $`W`$ such that $`|H_p||\delta _p|`$. 3. $`E_p`$ is a function which assigns a permutation $`e_{\pi ,\xi }^p\mathrm{Sym}(\{\xi \}\times \xi )`$ to each pair $`(\pi ,\xi )H_p\times \delta _p`$. We set $`q=(\delta _q,H_q,E_q)p=(\delta _p,H_p,E_p)`$ if and only if 1. $`\delta _p\delta _q`$; 2. $`H_pH_q`$; 3. $`E_pE_q`$; and 4. if $`\delta _p\xi <\delta _q`$, then the restriction to $`H_p`$ of the function, $`\pi e_{\pi ,\xi }^q`$, is an isomorphic embedding of $`H_p`$ into $`\mathrm{Sym}(\{\xi \}\times \xi )`$. ###### Claim 2.8. $``$ is $`\kappa `$-closed. ###### Proof of Claim 2.8. This is clear. ∎ ###### Claim 2.9. $``$ has the $`\kappa ^+`$-$`c.c.`$. ###### Proof of Claim 2.9. Suppose that $`p_i=(\delta _{p_i},H_{p_i},E_{p_i})`$ for $`i<\kappa ^+`$. Using the $`\mathrm{\Delta }`$-System Lemma and the assumption that $`\kappa ^{<\kappa }=\kappa `$, after passing to a suitable subsequence if necessary, we can suppose that the following conditions are satisfied. 1. There exists a fixed ordinal $`\delta `$ such that $`\delta _{p_i}=\delta `$ for all $`i<\kappa ^+`$. 2. There exists a fixed subgroup $`H`$ such that $`H_{p_i}H_{p_j}=H`$ for all $`i<j<\kappa ^+`$. 3. There exists a fixed function $`E`$ such that $`E_{p_i}H=E`$ for all $`i<\kappa ^+`$. Now fix any two ordinals $`i<j<\kappa ^+`$. Let $`H^+=H_{p_i},H_{p_j}`$ be the subgroup generated by $`H_{p_i}H_{p_j}`$; and let $`E^+`$ be any extension of $`E_{p_i}E_{p_j}`$ which satisfies condition (c). Then $`q=(\delta ,H^+,E^+)`$ is a common lower bound of $`p_i`$ and $`p_j`$. ∎ ###### Claim 2.10. For each $`\alpha <\kappa `$, the set $`C_\alpha =\{p\delta _p\alpha \}`$ is dense in $``$. ###### Proof of Claim 2.10. Let $`p=(\delta _p,H_p,E_p)`$. Then we can suppose that $`\delta _p<\alpha `$. We can define an isomorphic embedding $`\phi :H_p\mathrm{Sym}(H_p)`$ by setting $`\phi (h)(x)=hx`$ for all $`xH_p`$. Since $`\left|H_p\right|\left|\delta _p\right|`$, it follows that there exists an isomorphic embedding $`\phi _\xi :H_p\mathrm{Sym}(\{\xi \}\times \xi )`$ for each $`\delta _p\xi <\alpha `$. Hence there exists a condition $`q=(\delta _q,H_q,E_q)p`$ such that $`H_q=H_p`$ and $`\delta _q=\alpha `$. ∎ ###### Claim 2.11. For each $`\pi W`$, the set $`D_\pi =\{p\pi H_p\}`$ is dense in $``$. ###### Proof of Claim 2.11. Let $`p=(\delta _p,H_p,E_p)`$. Let $`H^+=H_p,\pi `$ be the subgroup generated by $`H_p\{\pi \}`$, and let $`E^+`$ be any extension of $`E_p`$ to $`H^+`$ which satisfies condition (c). Then $`q=(\delta _q,H_q,E_q)p`$. ∎ Let $`F`$ be a generic filter, and let $`V^{}=V[F]`$ be the corresponding generic extension. Working within $`V^{}`$, for each $`\pi W`$, let $$e(\pi )=\{e_{\pi ,\xi }^p\text{ There exists }pF\text{ such that }\pi H_p\text{ and }\xi <\delta _p\}.$$ Then $`e(\pi )\mathrm{Sym}(\mathrm{\Omega })`$. Let $`\mathrm{Sym}_\kappa (\mathrm{\Omega })=\{\psi \mathrm{Sym}(\mathrm{\Omega })|\mathrm{supp}(\psi )|<\kappa \}`$; and define the function $$f:W\mathrm{Sym}(\mathrm{\Omega })/\mathrm{Sym}_\kappa (\mathrm{\Omega })$$ by $`f(\pi )=e(\pi )\mathrm{Sym}_\kappa (\mathrm{\Omega })`$. Then it is enough to show that $`f`$ is an isomorphic embedding. ###### Claim 2.12. If $`1\pi W`$, then $`f(\pi )1`$. ###### Proof of Claim 2.12. Choose a condition $`p=(\delta _p,H_p,E_p)F`$ such that $`\pi H_p`$. If $`\xi `$ is any ordinal such that $`\delta _p\xi <\kappa `$, then $`e(\pi )\{\xi \}\times \xi id_{\{\xi \}\times \xi }`$. Hence $`e(\pi )\mathrm{Sym}_\kappa (\mathrm{\Omega })`$. ∎ ###### Claim 2.13. $`f`$ is a group homomorphism. ###### Proof of Claim 2.13. Let $`\pi _1`$, $`\pi _2W`$. Let $`p=(\delta _p,H_p,E_p)F`$ be a condition such that $`\pi _1`$, $`\pi _2H_p`$. Let $`\xi `$ be any ordinal such that $`\delta _p\xi <\kappa `$; and let $`qF`$ be a condition such that $`qp`$ and $`\xi <\delta _q`$. Then $$e_{\pi _1,\xi }^qe_{\pi _2,\xi }^q=e_{\pi _1\pi _2,\xi }^q.$$ It follows that $$e(\pi _1)\mathrm{Sym}_\kappa (\mathrm{\Omega })e(\pi _2)\mathrm{Sym}_\kappa (\mathrm{\Omega })=e(\pi _1\pi _2)\mathrm{Sym}_\kappa (\mathrm{\Omega }).$$ Finally it is easily checked that $`||=\lambda `$; and it follows that $`V^{}2^\kappa =\lambda `$. This completes the proof of Lemma 2.7. ∎ Summing up our work in this section, we obtain the following result. ###### Theorem 2.14. Let $`V`$ be a transitive model of $`ZFC`$ and let $`\kappa `$, $`\lambda `$ be cardinals such that $`\omega <\kappa ^{<\kappa }=\kappa <\lambda =\lambda ^\kappa `$. Let $`\alpha `$ be any ordinal such that $`\alpha <\lambda ^+`$. Then there exists a notion of forcing $``$ such that 1. $``$ is $`\kappa `$-closed; 2. $``$ has the $`\kappa ^+`$-$`c.c.`$; and such that the following statements are true in the generic extension $`V^{}`$. 1. $`2^\kappa =\lambda `$. 2. There exist groups $`H<G<\mathrm{Sym}(\kappa )`$ such that $`|H|=\kappa `$ and the normaliser tower of $`H`$ in $`G`$ terminates in exactly $`\alpha `$ steps. ###### Proof. Let $`\gamma `$ be an ordinal such that $`\mathrm{max}\{\alpha ,\lambda \}\gamma <\lambda ^+`$. Then $`|W_\gamma |=\lambda `$. Let $``$ be the notion of forcing obtained by applying Lemma 2.7 to $`W=W_\gamma `$. By Lemma 2.5, $``$ satisfies our requirements. ∎ ## 3. Closed groups of uncountable degree In this section, we will prove Theorem 1.12. Let $`\kappa `$ be a regular uncountable cardinal such that $`\kappa ^{<\kappa }=\kappa `$, and let $`G`$ be a group which satisfies the $`\kappa ^+`$-compatibility condition. Let $`L`$ be a first-order language consisting of $`\kappa `$ binary relation symbols. The following notion of forcing $``$ is designed to adjoin a structure $``$ of cardinality $`\kappa `$ for the language $`L`$ such that $`G\mathrm{Aut}`$. This is sufficient; for then we can use one of the standard procedures to code $``$ into a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`\mathrm{Aut}\mathrm{\Gamma }\mathrm{Aut}`$. ( Cf. Section 5.5 of Hodges .) ###### Definition 3.1. Suppose that $`L_0L`$ and that $`𝒩`$ is a structure for the language $`L_0`$. Then a restricted atomic type in the free variable $`v`$ for the language $`L_0`$ using parameters from $`𝒩`$ is a set $`t`$ of formulas of the form $`R(v,a)`$, where $`RL_0`$ and $`a𝒩`$. An element $`c𝒩`$ is said to realise $`t`$ if $`𝒩\phi [c]`$ for every formula $`\phi (v)t`$. If no element of $`𝒩`$ realises $`t`$, then $`t`$ is said to be omitted in $`𝒩`$. Notice that if $`t`$ is omitted in $`𝒩`$, then $`t`$ is not the trivial restricted atomic type $`\mathrm{}`$. ###### Definition 3.2. Let $``$ be the notion of forcing consisting of the conditions $$p=(H,\pi ,𝒩,T)$$ such that the following hold. 1. $`H`$ is a subgroup of $`G`$ such that $`\left|H\right|<\kappa `$. 2. There exists an ordinal $`0<\delta <\kappa `$ and a subset $`L(𝒩)[L]^{<\kappa }`$ such that $`𝒩`$ is a structure with universe $`\delta `$ for the language $`L(𝒩)`$. 3. $`\pi :H\mathrm{Aut}𝒩`$ is a group homomorphism. 4. $`T`$ is a set of restricted atomic types in the free variable $`v`$ for the language $`L(𝒩)`$ using parameters from $`𝒩`$. Furthermore, $`\left|T\right|<\kappa `$; and each $`tT`$ is omitted in $`𝒩`$. We set $`(H_2,\pi _2,𝒩_2,T_2)(H_1,\pi _1,𝒩_1,T_1)`$ if and only if 1. $`H_1H_2`$. 2. $`𝒩_1`$ is a substructure of $`𝒩_2`$. 3. For all $`hH_1`$ and $`\alpha 𝒩_1`$, $`\pi _2(h)(\alpha )=\pi _1(h)(\alpha )`$. 4. $`T_1T_2`$. It is clear that the components $`(H,\pi ,𝒩)`$ in each condition $`p`$ are designed to generically adjoin a structure $``$ of cardinality $`\kappa `$ for the language $`L`$, together with an embedding $`\pi ^{}`$ of $`G`$ into $`\mathrm{Aut}`$. The set $`T`$ of restricted atomic types is needed to kill off potential extra automorphisms $`g\mathrm{Aut}\pi ^{}[G]`$, and thus ensure that $`\pi ^{}`$ is surjective. ###### Lemma 3.3. $``$ is $`\kappa `$-closed. ###### Proof. This is clear. ∎ ###### Lemma 3.4. $``$ has the $`\kappa ^+`$-$`c.c.`$ ###### Proof. Suppose that $`p_i=(H_i,\pi _i,𝒩_i,T_i)`$ for $`i<\kappa ^+`$. After passing to a suitable subsequence if necessary, we can suppose that the following conditions hold. 1. There exists a fixed structure $`𝒩`$ such that $`𝒩_i=𝒩`$ for all $`i<\kappa ^+`$. 2. There exists a fixed set of restricted atomic types $`T`$ such that $`T_i=T`$ for all $`i<\kappa ^+`$. 3. There is a fixed group $`H`$ such that for each $`i<\kappa ^+`$, there exists an isomorphism $`f_i:HH_i`$. 4. For each $`i<\kappa ^+`$, let $`\psi _i:H\mathrm{Aut}𝒩`$ be the embedding defined by $`\psi _i=\pi _if_i`$. Then $`\psi _i=\psi _j`$ for all $`i<j<\kappa ^+`$. Since $`G`$ satisfies the $`\kappa ^+`$-compatibility condition, there exist ordinals $`i<j<\kappa ^+`$ and a surjective homomorphism $`\phi :H_i,H_jH_i`$ such that 1. $`\phi f_j=f_i`$; and 2. $`\phi H_i=id_{H_i}`$. Let $`H_i,H_j\mathrm{Aut}𝒩`$ be the homomorphism defined by $`\pi =\pi _i\phi `$. Clearly $`\pi _i\pi `$. Note that if $`xH_j`$, then $`\pi _i\phi (x)`$ $`=\pi _i\left(\phi f_j\right)f_j^1(x)`$ $`=\pi _if_if_j^1(x)`$ $`=\pi _j(x).`$ Thus we also have that $`\pi _j\pi `$. Consequently, we can define a condition $`pp_i`$, $`p_j`$ by $$p=(H_i,H_j,\pi ,𝒩,T).$$ ###### Lemma 3.5. For each $`p=(H,\pi ,𝒩,T)`$, there exists $`p^+=(H,\pi ^+,𝒩^+,T)p`$ such that $`\pi ^+:H\mathrm{Aut}𝒩^+`$ is an embedding. ###### Proof. Let $`𝒩^+`$ be the structure for the language $`L(𝒩)`$ such that 1. the universe of $`𝒩^+`$ is the disjoint union $`𝒩H`$; 2. for each relation $`RL(𝒩)`$, $`R^{𝒩^+}=R^𝒩`$. In particular, if $`x𝒩^+𝒩`$, then $`x`$ only realises the trivial restricted atomic type $`\mathrm{}`$ over $`𝒩`$. Hence none of the restricted atomic types in $`T`$ is realised in $`𝒩^+`$. Let $`\pi ^+:H\mathrm{Aut}𝒩^+`$ be the embedding such that for each $`hH`$, 1. $`\pi ^+(h)(x)=\pi (h)(x)`$ for all $`x𝒩`$; and 2. $`\pi ^+(h)(x)=hx`$ for all $`xH`$. Then $`p^+=(H,\pi ^+,𝒩^+,T)p`$. ∎ There is a slight inaccuracy in the proof of Lemma 3.5, as the universe of $`𝒩^+`$ should really be an ordinal $`\delta <\kappa `$. However, the proof can easily be repaired: simply replace $`𝒩^+`$ by a suitable isomorphic structure. Similar remarks apply to the proofs of Lemmas 3.6 and 3.8. ###### Lemma 3.6. For each $`aG`$, $$D_a=\{(H,\pi ,𝒩,T)aH\}$$ is a dense subset of $``$. ###### Proof. Let $`aG`$ and $`p=(H,\pi ,𝒩,T)`$. We can suppose that $`aH`$. Let $`H^+=H,a`$. Let $`C=\{g_iiI\}`$ be a set of left coset representatives for $`H`$ in $`H^+`$, chosen so that $`1C`$. Let $`𝒩^+`$ be the structure for the language $`L(𝒩)`$ such that 1. the universe of $`𝒩^+`$ is the cartesian product $`C\times 𝒩`$; and 2. for each relation $`RL(𝒩)`$, $$((g_i,x),(g_j,y))R^{𝒩^+}\text{ iff }i=j\text{ and }(x,y)R^𝒩.$$ By identifying each $`x𝒩`$ with the element $`(1,x)𝒩^+`$, we can regard $`𝒩`$ as a substructure of $`𝒩^+`$. Once again, each element $`(g_i,x)𝒩^+𝒩`$ only realises the trivial restricted atomic type $`\mathrm{}`$ over $`𝒩`$; and hence none of the restricted atomic types in $`T`$ is realised in $`𝒩^+`$. Define an action of $`H^+`$ on $`𝒩^+`$ as follows. If $`gH^+`$ and $`(g_i,x)𝒩^+`$, then $$g(g_i,x)=(g_j,\pi (h)(x)),$$ where $`jI`$ and $`hH`$ are such that $`gg_i=g_jh`$. It is easily checked that this action yields a homomorphism $`\pi ^+:H^+\mathrm{Aut}𝒩^+`$; and that $`(H^+,\pi ^+,𝒩^+,T)p`$. ∎ ###### Lemma 3.7. For each $`\alpha <\kappa `$, $$E_\alpha =\{(H,\pi ,𝒩,T)\alpha 𝒩\}$$ is a dense subset of $``$. ###### Proof. Left to the reader. ∎ Let $`F`$ be a generic filter, and let $`V^{}=V[F]`$ be the corresponding generic extension. Working within $`V^{}`$, define $$=\{𝒩\text{ There exists }p=(H,\pi ,𝒩,T)F\}$$ and $$\pi ^{}=\{\pi \text{ There exists }p=(H,\pi ,𝒩,T)F\}.$$ Then $``$ is a structure for $`L`$ of cardinality $`\kappa `$, and $`\pi ^{}`$ is an embedding of $`G`$ into $`\mathrm{Aut}`$. So the following lemma completes the proof of Theorem 1.12. ###### Lemma 3.8. $`\pi ^{}:G\mathrm{Aut}`$ is a surjective homomorphism. ###### Proof. Suppose that $`g\mathrm{Aut}\pi ^{}[G]`$. Let $`\stackrel{~}{}`$, $`\stackrel{~}{\pi }`$ be the canonical $``$-names for $``$ and $`\pi ^{}`$; and let $`\stackrel{~}{g}`$ be a $``$-name for $`g`$. Then there exists a condition $`pF`$ such that $$p\stackrel{~}{g}\mathrm{Aut}\stackrel{~}{}\text{ and }\stackrel{~}{g}\stackrel{~}{\pi }(h)\text{ for all }hG.$$ Using the fact that $``$ is $`\kappa `$-closed, we can inductively construct a descending sequence of conditions $`p_m=(H_m,\pi _m,𝒩_m,T_m)`$ for $`m\omega `$ such that the following hold. 1. $`p_0=p`$. 2. For all $`x𝒩_m`$, there exists $`y𝒩_{m+1}`$ such that $`p_{m+1}\stackrel{~}{g}(x)=y`$. 3. For all $`hH_m`$, there exists $`z𝒩_{m+1}`$ such that $`p_{m+1}\stackrel{~}{g}(z)\pi _{m+1}(h)(z)`$. Let $`q=(H,\pi ,𝒩,T)`$ be the greatest lower bound of $`\{p_mm\omega \}`$ in $``$. Then $`qp`$, and there exists $`g^{}\mathrm{Aut}𝒩\pi [H]`$ such that $`q\stackrel{~}{g}𝒩=g^{}`$. Let $`𝒩^+`$ be the structure defined as follows. 1. The universe of $`𝒩^+`$ is the disjoint union $`𝒩H`$. 2. For each relation $`RL(𝒩)`$, $`R^{𝒩^+}=R^𝒩`$. 3. For each $`x𝒩`$, let $`R_xLL(𝒩)`$ be a new binary relation symbol. Then we set $`(h,y)R_x^{𝒩^+}`$ iff $`hH`$, $`y𝒩`$ and $`\pi (h)(x)=y`$. Once again, it is clear that none of the restricted atomic types in $`T`$ is realised in $`𝒩^+`$. Let $`\pi ^+:H\mathrm{Sym}𝒩^+`$ be the embedding such that 1. $`\pi ^+(h)(x)=\pi (h)(x)`$ for all $`x𝒩`$; and 2. $`\pi ^+(h)(x)=hx`$ for all $`xH`$. Then it is easily checked that $`\pi ^+[H]\mathrm{Aut}𝒩^+`$. Finally let $`t`$ be the restricted atomic type defined by $$t=\{R_x(v,g^{}(x)x𝒩\};$$ and let $`T^+=T\{t\}`$. ###### Claim 3.9. $`t`$ is omitted in $`𝒩^+`$. ###### Proof of Claim 3.9. If $`z𝒩^+`$ realises $`t`$, then $`z=hH`$. And if $`x𝒩`$, then $`𝒩^+R_x(h,g^{}(x))`$, and so $`\pi (h)(x)=g^{}(x)`$. But this contradicts the fact that $`g^{}\mathrm{Aut}𝒩\pi [H]`$. ∎ Thus $`q^+=(H,\pi ^+,𝒩^+,T^+)`$. To simplify notation, suppose that $`q^+F`$; so that $`g^{}g`$. Then for each $`x𝒩`$, we have that $`R_x(1,x)`$, and hence $`R_x(g(1),g^{}(x))`$. But this means that $`g(1)`$ realises $`t`$, which is the final contradiction. ∎ ## 4. $`\tau _\kappa `$ is increasing In this section, we will complete the proof of Theorem 1.4. So let $`VGCH`$ and let $`\kappa `$, $`\lambda V`$ be uncountable cardinals such that $`\kappa <\mathrm{cf}(\lambda )`$. Let $`\alpha `$ be any ordinal such that $`\alpha <\lambda ^+`$. It is well-known that there exists a centreless group $`T`$ of cardinality $`\kappa `$ such that $`\mathrm{Aut}T=\mathrm{Inn}T`$. (For example, we can take $`T=PGL(2,K)`$, where $`K`$ is a rigid field of cardinality $`\kappa `$. The existence of such a field follows from Lemma 1.7.) Thus we can assume that $`\alpha 1`$. First consider the case when $`\kappa `$ is a regular cardinal. Let $``$ be the notion of forcing which is given by Theorem 2.14. Then the following statements are true in the corresponding generic extension $`M=V^{}`$. 1. $`\kappa ^{<\kappa }=\kappa `$. 2. $`2^\kappa =\lambda `$. 3. There exist groups $`H<G<\mathrm{Sym}(\kappa )`$ such that $`|H|=\kappa `$ and the normaliser tower of $`H`$ in $`G`$ terminates in exactly $`\alpha `$ steps. Let $`M`$ be the notion of forcing, given by Theorem 1.12, which adjoins a graph $`\mathrm{\Gamma }`$ of cardinality $`\kappa `$ such that $`G\mathrm{Aut}\mathrm{\Gamma }`$. Then, applying Lemma 1.8, we find that the following statements are true in $`M^{}`$. 1. $`2^\kappa =\lambda `$. 2. There exists a centreless group $`T`$ of cardinality $`\kappa `$ such that $`\tau (T)=\alpha `$. Next suppose that $`\kappa `$ is a singular cardinal. By the above argument, there is a generic extension $`V^{}`$ in which the following statements are true. 1. $`2^{\omega _1}=2^\kappa =\lambda `$. 2. There exists a centreless group $`T`$ of cardinality $`\omega _1`$ such that $`\tau (T)=\alpha `$. Let $`G=T\times \mathrm{Alt}(\kappa )`$. Clearly $`|G|=\kappa `$; and the following theorem implies that $`\tau (G)=\tau (T)=\alpha `$. This completes the proof of Theorem 1.4. ###### Theorem 4.1. Suppose that $`\omega \theta <\kappa `$. If $`H`$ is a centreless group of cardinality $`\theta `$ such that $`\tau (H)1`$, then $`\tau (H\times \mathrm{Alt}(\kappa ))=\tau (H)`$. ###### Corollary 4.2. If $`\omega \theta <\kappa `$, then $`\tau _\theta \tau _\kappa `$. The remainder of this section will be devoted to the proof of Theorem 4.1. Let $`G=H\times \mathrm{Alt}(\kappa )`$; and let $`H_\beta `$, $`G_\beta `$ be the $`\beta ^{th}`$ groups in the automorphism towers of $`H`$, $`G`$ respectively. We will eventually prove by induction that $`G_\beta =H_\beta \times \mathrm{Sym}(\kappa )`$ for all $`\beta 1`$. To accomplish this, we need to keep track of $`\phi [\mathrm{Alt}(\kappa )]`$ for each automorphism $`\phi `$ of $`G_\beta `$. The next lemma shows that for all $`\phi \mathrm{Aut}G_\beta `$, either $`\phi [\mathrm{Alt}(\kappa )]H_\beta `$ or $`\phi [\mathrm{Alt}(\kappa )]\mathrm{Sym}(\kappa )`$. The main point will be to eliminate the possibility that $`\phi [\mathrm{Alt}(\kappa )]H_\beta `$. This is straightforward when $`\beta `$ is a successor ordinal. To deal with the case when $`\beta `$ is a limit ordinal, we will make use of the result that $`\mathrm{Alt}(\kappa )`$ is strictly simple. ###### Lemma 4.3. Suppose that $`A`$ is a simple nonabelian normal subgroup of the direct product $`H\times S`$. Then either $`AH`$ or $`AS`$. ###### Proof. Let $`1g=xyA`$, where $`xH`$ and $`yS`$. If $`y=1`$, then the conjugacy class $`g^A=x^A`$ is contained in $`H`$, and so $`A=g^AH`$. So suppose that $`y1`$. Let $`\pi :H\times SS`$ be the canonical projection map. Then $`1y\pi [A]S`$ and $`\pi [A]A`$. Hence there exists an element $`z\pi [A]S`$ such that $`zyz^1y`$. Since $`A\mathrm{}H\times S`$, it follows that $$1zyz^1y^1=zxyz^1y^1x^1=zgz^1g^1AS.$$ Arguing as above, we now obtain that $`AS`$. ∎ ###### Definition 4.4. Let $`H`$ be a subgroup of the group $`G`$. Then $`H`$ is said to be an ascendant subgroup of $`G`$ if there exist an ordinal $`\gamma `$ and a set of subgroups $`\{H_\beta \beta \gamma \}`$ such that the following conditions are satisfied. 1. $`H_0=H`$ and $`H_\gamma =G`$. 2. If $`\beta <\gamma `$, then $`H_\beta \mathrm{}H_{\beta +1}`$. 3. If $`\delta `$ is a limit ordinal such that $`\delta \gamma `$, then $`H_\delta =_{\beta <\delta }H_\beta `$. ###### Definition 4.5. A group $`A`$ is strictly simple if it has no nontrivial proper ascendant subgroups. ###### Theorem 4.6 (Macpherson and Neumann ). For each $`\kappa \omega `$, the alternating group $`\mathrm{Alt}(\kappa )`$ is strictly simple. $`\mathrm{}`$ ###### Lemma 4.7. Suppose that $`H`$ is a centreless group. Let $`\tau =\tau (H)`$ and let $`H_\tau `$ be the $`\tau ^{th}`$ group in the automorphism tower of $`H`$. If $`A`$ is a strictly simple normal subgroup of $`H_\tau `$, then $`AH_0`$. ###### Proof. Let $`\beta \tau `$ be the least ordinal such that $`AH_\beta `$. First suppose that $`\beta `$ is a limit ordinal. Then $`A=_{\gamma <\beta }\left(AH_\gamma \right)`$, and $`AH_\gamma \mathrm{}AH_{\gamma +1}`$ for all $`\gamma <\beta `$. Consequently, if $`\gamma <\beta `$ is the least ordinal such that $`AH_\gamma 1`$, then $`AH_\gamma `$ is a nontrivial ascendant subgroup of $`A`$. But this contradicts the assumption that $`A`$ is strictly simple. Next suppose that $`\beta =\gamma +1`$ is a successor ordinal. Since $`AH_\gamma `$ is a proper normal subgroup of $`A`$, it follows that $`AH_\gamma =1`$. Now notice that $`AH_{\gamma +1}N_{H_\tau }(H_\gamma )`$ and $`H_\gamma H_\tau =N_{H_\tau }(A)`$, This implies that $`[A,H_\gamma ]AH_\gamma =1`$. (For example, see Lemma 1.1.3 of Suzuki .) But then $`AC_{H_{\gamma +1}}(H_\gamma )=1`$, which is a contradiction. The only remaining possibility is that $`\beta =0`$. ∎ We will also make use of the well-known results that $`\mathrm{Aut}(\mathrm{Alt}(\kappa ))=\mathrm{Sym}(\kappa )`$ and $`\mathrm{Aut}(\mathrm{Sym}(\kappa ))=\mathrm{Sym}(\kappa )`$. (For example, see Theorem 11.4.8 of Scott .) ###### Proof of Theorem 4.1. Let $`\tau =\tau (H)`$; and let $$H=H_0\mathrm{}H_1\mathrm{}\mathrm{}H_\beta \mathrm{}H_{\beta +1}\mathrm{}\mathrm{}H_\tau =H_{\tau +1}=\mathrm{}$$ be the automorphism tower of of $`H`$. Let $`G=H\times \mathrm{Alt}(\kappa )`$; and let $$G=G_0\mathrm{}G_1\mathrm{}G_2\mathrm{}\mathrm{}G_\beta \mathrm{}G_{\beta +1}\mathrm{}\mathrm{}$$ be the automorphism tower of $`G`$. We will prove by induction on $`\beta 1`$ that $`G_\beta =H_\beta \times \mathrm{Sym}(\kappa )`$. First consider the case when $`\beta =1`$. Let $`\phi \mathrm{Aut}G`$ be any automorphism. Then $`\phi [\mathrm{Alt}(\kappa )]`$ is a simple nonabelian normal subgroup of the direct product $`G=H\times \mathrm{Alt}(\kappa )`$. By Lemma 4.3, either $`\phi [\mathrm{Alt}(\kappa )]H`$ or $`\phi [\mathrm{Alt}(\kappa )]\mathrm{Alt}(\kappa )`$. Since $`\left|\phi [\mathrm{Alt}(\kappa )]\right|=\kappa >\theta =|H|`$, it follows that $`\phi [\mathrm{Alt}(\kappa )]\mathrm{Alt}(\kappa )`$. As $`\phi [\mathrm{Alt}(\kappa )]`$ is a normal subgroup of $`G`$, we must have that $`\phi [\mathrm{Alt}(\kappa )]=\mathrm{Alt}(\kappa )`$. Note that $`C_G(\mathrm{Alt}(\kappa ))=H`$. Hence we must also have that $`\phi [H]=H`$. It follows that $$G_1=\mathrm{Aut}G=\mathrm{Aut}H\times \mathrm{Aut}(\mathrm{Alt}(\kappa ))=H_1\times \mathrm{Sym}(\kappa ).$$ Next suppose that $`\beta =\gamma +1`$ and that $`G_\gamma =H_\gamma \times \mathrm{Sym}(\kappa )`$. Let $`\phi \mathrm{Aut}G_\gamma `$ be any automorphism. By Lemma 4.3, either $`\phi [\mathrm{Alt}(\kappa )]H_\gamma `$ or $`\phi [\mathrm{Alt}(\kappa )]\mathrm{Sym}(\kappa )`$. As $`\mathrm{Alt}(\kappa )`$ is a strictly simple group, Lemma 4.7 implies that $`\phi [\mathrm{Alt}(\kappa )]\mathrm{Sym}(\kappa )`$. Since $`\phi [\mathrm{Alt}(\kappa )]`$ is a simple normal subgroup of $`G_\gamma `$, it follows that $`\phi [\mathrm{Alt}(\kappa )]=\mathrm{Alt}(\kappa )`$. Using the facts that $`C_{G_\gamma }(\mathrm{Alt}(\kappa ))=H_\gamma `$ and $`C_{G_\gamma }(H_\gamma )=\mathrm{Sym}(\kappa )`$, we now see that $`\phi [H_\gamma ]=H_\gamma `$ and $`\phi [\mathrm{Sym}(\kappa )]=\mathrm{Sym}(\kappa )`$. Hence $$G_{\gamma +1}=\mathrm{Aut}G_\gamma =\mathrm{Aut}H_\gamma \times \mathrm{Aut}(\mathrm{Sym}(\kappa ))=H_{\gamma +1}\times \mathrm{Sym}(\kappa ).$$ Finally no difficulties arise when $`\beta `$ is a limit ordinal. ∎
warning/0003/astro-ph0003137.html
ar5iv
text
# The CFH Optical PDCS survey (COP) I: The Data ## 1 Introduction One of the main goals of the study of distant rich clusters of galaxies is to understand their origin and evolution. Clusters are invaluable cosmological probes, since the evolution of cluster abundances is strongly dependent on the underlying cosmology and therefore can constrain cosmological models (e.g. Bahcall et al. 1997, Oukbir & Blanchard 1992 and 1997, Reichart et al 1999, Nichol et al. 1999). In order to be able to exploit this potential, large statistically representative spectroscopic samples are, however, required. There are two types of samples that are being developed: those based on optical selection criteria and those based on X-ray detections. Below z of about 0.1, optically selected studies (e.g. ENACS: Katgert et al. 1996) have had about as many clusters in them as those selected via their X-ray fluxes. Up until recently, however, the higher redshift work has been dominated by X-ray selection techniques (e.g. CNOC: Carlberg et al. 1996 SHARC: Romer et al. 2000; RDCS: Rosati et al. 1998; WARPS: Jones et al. 1998; Vikhlinin et al. 1998). To greatly enlarge the sample of detailed studies of redshift about 0.4 optically selected clusters, we have embarked on a photometric and redshift campaign based on the Palomar Distant Cluster survey (Postman et al. 1996, see also Holden et al. 1999). We have observed a significant number of regions on the sky (10) and obtained about 70 redshifts per line of sight. These pointings were known to contain candidate clusters of galaxies based on the PDCS studies (e.g. Postman et al. 1996, Holden et al. 1997). The main purpose of this paper is to publish the COP survey data and to describe the data reduction so as to lay a foundation for future papers. The interpretation of the results is, therefore, given in later papers (e.g. Holden et al. 2000). The outline of this paper is as follows. In Section 2, we give the observational strategy. In Section 3, we describe the way we have reduced and analyzed our photometry. In Section 4, we describe the way we have reduced and analyzed our spectroscopy. In the last section, we give an analysis of the redshift and spatial distribution of the galaxies in our sample. The data are given in tables 6-15. ## 2 Target Selection and Observations ### 2.1 Observing strategy Our project required the measurement of a large number of redshifts ($``$ 100) of faint galaxies ($`V_{PDCS}`$$``$23) for a significant number of clusters ($``$ 10). It was necessary, therefore, to optimize our spectroscopic observations to get as many useful spectra as possible per night. We had photometric data in our fields (the PDCS catalog: Postman et al. 1996) prior to our CFH observations which considerably reduced the number of nights necessary to produce the spectroscopic catalogue (compared for example to the time needed to achieve the CNOC survey, Yee et al. 1996). The first goal of the survey was to study the reality of the selected cluster candidates: are these real physical systems or are these only galaxy number count enhancements due to superposition effects (see Holden et al. 2000)? This placed a requirement on the number of redshifts we needed along the line of sight (Katgert et al. 1996). Moreover, we wanted to compute a global velocity dispersion for each cluster. This required $``$10 redshifts in the main groups (the $`clusters`$) to allow us to use robust estimators (see e.g. Adami et al. 1998c). Assuming a line-of-sight contamination between 50 and 75% for a cluster at z$``$0.4 (see e.g. Carlberg et al. 1996) and a success rate of 70% (see Adami et al. 1998b) for our magnitude ranges, we needed to obtain between 50 and 60 spectra for each line of sight in the ideal situation of 1 cluster per line of sight. Since there could be 2 structures (or more) per pointing, however, we set a goal of measuring 100 spectra for each pointing (to yield about 70 redshifts). In order to have a statistically representative set of more than 10 lines of sight with more than 70 redshifts, we have used the CFH-MOS multi-slits spectrograph for its high multiplex gain. We wanted to measure the radial velocity of our targets with an uncertainty of less than 150 km s<sup>-1</sup> because this precision is almost the same as the one obtained for ENACS and CNOC galaxies (e.g. Katgert et al 1996, Mazure et al 1996, Yee et al. 1996). This allows an accurate comparison with these two surveys. Following Adami et al. (1998b) and Yee et al. (1996), we have used the CFH O300 grism, which provides a dispersion of about 5 Å.px<sup>-1</sup> with the STIS2 CFH CCD (pixels of 0.43”). The precision in the velocity measurement depends of both the resolution given by the grism and slit width (theoretical limiting factor) and the observational conditions (observational limiting factor). The resolution of the O300 grism allowed us to reach the required velocity accuracy. We observed extended objects that were a few arcsecs in diameter. In order to properly subtract the sky in our spectra, we used slitlet lengths of 11 arcsecs. This setup allowed us to place about 40 slitlets per mask for the full spectral range of about 6500 Å delivered by the grism to be used. The setup would require about 3 masks per pointing to reach our goal of 100, however. In this case, the amount of time needed to observe 10 line of sights would have been prohibitive. To reduce the amount of time required and increase the multiplexing capabilities of the instrument to place $``$70 slitlets on each mask, we used CFH blocking filters (see Table 1 and see also the CNOC survey: Yee et al. 1996). This was effective, since we had estimates of the cluster redshifts (Postman et al. 1996), which have been confirmed to be accurate (see Holden et al. 1997) enough for our purposes. These estimates allowed us to chose the right CFH filter so as to span a spectral range that included at least 3 lines for each spectrum (typically selected from \[OII\], H&K, G band, H$`\beta `$, \[OIII\] and H$`\alpha `$). For each of the two filters used, Table 1 shows the redshift range that gives dispersed spectra of galaxies at that redshift that include the \[OII\], H&K and G band spectral features. Table 2 gives the filter used for each line of sight. ### 2.2 Cluster candidate selection We selected cluster candidates to match the CFH telescope capabilities. It was impossible with the telescope time availability to sample structures at redshifts significantly greater than 0.5 (see Lubin et al. 1998 and references therein for such a study). We decided, therefore, to restrict our sample to cluster candidates in the estimated redshift range $`0.3<z<0.5`$. We also selected clusters so as to be able to complete the sample in only two semesters at the CFHT. Therefore, targets were selected from the PDCS fields at 9 and 13 hours for the Spring semester and from the PDCS fields at 16, 0 and 2 hours for the Autumn semester. We selected also the candidate clusters with both a richness class 1 or greater and significant density peaks in the galaxy distribution (see Fig. 1, 2 and 3) and we used the highest galaxy density areas. These densest areas coincided with or were close to the cluster centers given in Lubin et al. (1996) in most cases. PDCS34 was the exception. We observed at a position about 5’ to the North of the given cluster center, slightly different from Lubin et al. (1996), as we found no galaxy concentration at the exact coordinates of PDCS34. In order to describe the galaxy distribution on the sky, we have computed the local projected galaxy density and produced isodensity contours for the PDCS galaxies using an adaptative kernel technique (e.g. Adami et al 1998a and ref. therein) for each line of sight. This technique adapts the size of the map window to the local density of objects. The same technique has been used in Adami et al (1998a) to study the ENACS clusters. Where the galaxy density is higher, the window used to compute the density of objects is reduced to a value consistent with producing a statistically significant number of galaxies. Where the galaxy density is lower, the window is larger so as to produce a similarly valid statistic for the density estimate. We have then produced the Figures 1,2 and 3. Finally, among the cluster candidates matching the previous conditions, we selected those ones that were detected in X-rays (Holden et al. 1997, 1999) whenever possible (only 3 of the 10 lines of sight). ### 2.3 Mask design We optimized the number of slitlets per mask to increase the efficiency of the survey. It is possible to show that for a field with a very high density of targets, the optimal configuration is to place the slitlets in band configurations. With our limiting magnitude and slitlet width, however, this is not the best method to use because the density of targets is not always the same. To take this into account, we have adapted the Minimal Spanning Tree (MST hereafter) method (e.g. Dussert et al. 1986) in order to find the optimal configuration according to the filter used. We only give a brief description here: for a given set of points, the MST process finds the minimal total length of a tree covering this set (without a loop). If we fix the area (thus the length of tree) the MST exactly finds the maximum number of slits that can be put in that area (according to the constraints: size of the slits, magnitudes, filter ..etc…). We have checked this method by showing that it gives a band configuration for a high density of targets. We show in Figures 4 and 5 the typical configuration given by this method for 2 of the lines of sight: PDCS62 and PDCS67. PDCS62 has a high density of targets (3.97 gal arcmin<sup>2</sup>) while PDCS67 has a lower target density (2.20 gal arcmin<sup>2</sup>). The slit distribution for PDCS62 is close to a band configuration. Practically, we designed the masks in three steps. First, the primary potential targets were selected from the galaxies with a low enough magnitude to provide a reasonable success rate (percentage of observed galaxies with a redshift successfully measured) according to the planned exposure time. This exposure time was chosen to observe galaxies brighter than V$`{}_{PDCS}{}^{}19`$ at the mean redshift of the survey (z$``$0.4). This is about 1.5 magnitude fainter than the typical values of M in nearby clusters (see e.g. Rauzy et al. 1998). We used the MST selection for these galaxies first. Table 2 gives these magnitude ranges with the real success rates. The mean value is 66$`\%`$, only slightly lower than the expected value. Then, the secondary potential targets were selected from the galaxies in the next 0.5 magnitude bin (in principle too faint to provide the same S/N, see Table 2 for the $`V_{PDCS}`$ magnitude range). Finally, if some space remained on the mask after selecting these two types of targets, we also assigned slitlets to contain other objects, typically in the $`V_{PDCS}`$ magnitude range or selected by ”eye” during the night (tertiary targets). We did not select the galaxies on the basis of their color, in order to avoid selection effects along the line of sight. Also, the second mask for PDCS62 has been partially designed by hand (during the night with the image acquired at the CFH telescope) because the PDCS photometric data did not cover all the field (only about 50$`\%`$). ## 3 The photometry ### 3.1 The PDCS data We selected 10 lines of sight (see Table 2 and Fig. 1-3 and 6) for our observations which include 14 PDCS cluster candidates that are described in Postman et al. (1996). We have used the original PDCS photometry to select the galaxy targets. The PDCS photometry was carried out in the 4-shooter Palomar $`V`$ and $`I`$ filters and calibrated in the AB system. We will refer to it as $`V_{PDCS}`$ and $`I_{PDCS}`$ from now on. Postman et al. (1996) showed how this photometry compared to the Vega-system standard system. Here, we briefly describe the comparison between systems: the effective wavelength of the $`V_{PDCS}`$ filter is $``$100Å bluer and about 50$`\%`$ wider than the standard Johnson’s $`V`$. $`I_{PDCS}`$ has nearly the same width as the Kron-Cousins $`I`$ filter, but the effective wavelength is about 500Å redder. The zero points of the $`V_{PDCS}`$ and $`I_{PDCS}`$ magnitudes are based on the $`AB`$ magnitude system of Oke $`\&`$ Gunn (1983). The magnitudes of Vega are $`V_{PDCS}`$=+0.03 and $`I_{PDCS}`$=+0.46. The relation between ($`V_{PDCS}`$, $`I_{PDCS}`$) and (V,I) are: $`V=V_{PDCS}0.020.056(V_{PDCS}I_{PDCS})+0.012(V_{PDCS}I_{PDCS})^2`$ and $`I=I_{PDCS}0.43+0.089(V_{PDCS}I_{PDCS})`$ The uncertainty for $`V_{PDCS}I_{PDCS}`$ is almost 0.2 magnitude (Lubin 1996). Since we used the PDCS star/galaxy classification to select the galaxies to observe, it is of interest to determine how well this selection performed. Given our observational strategy with blocking filters, faint stars remain unidentified because no obvious spectral feature falls within our spectral coverage. The same is true for faint galaxies at redshifts outside the optimal range of the filter used and for galaxies only detected at low signal-to-noise. The validity of the star/galaxy separation can only then be tested against other methods for these cases. We have thus compared the PDCS star/galaxy selection against the classification scheme of Sextractor (Bertin $`\&`$ Arnouts 1996). Figure 7 shows the comparison for the West PDCS62 spectroscopic field using a $`V`$ image taken at CFH (see below). We have plotted the ANN (Artificial Neural Network) parameter of Sextractor characterizing the nature of the objects versus the $`V_{PDCS}`$ magnitude for the PDCS objects classified as $`galaxies`$. The objects in Figure 7 are, therefore, only galaxies according to the PDCS classification. The Sextractor ANN parameter spans the range , being close to 1 if the object is classified by Sextractor as a star and moving closer to 0 if the object resembles a galaxy. There is a low contamination rate at faint magnitudes: a few number of objects classified as galaxies by the PDCS are interpreted as stars if we use Sextractor. We conclude, therefore, that we optimized the selection of our targets as can be seen by the ”ridge line” near 0 in Figure 7. However, according to the mask design technique, we targeted sometimes, in the tertiary-target-class, objects that were classified as stars by the PDCS, but which were later found to be galaxies. These objects represent, however, less than 3$`\%`$ of the sample and are not a significant source of error. ### 3.2 The CFH photometry Besides the multislit spectroscopy, we have also imaged the fields of study with the CFHT. The field of view of the frames obtained was $`10^{}\times 10^{}`$. The imaged areas are shown in Fig. 6 (the photometric fields are slightly larger than the spectroscopic fields of Fig. 6 which cover about $`8^{}\times 8^{}`$). We used the $`R`$, $`V`$, $`2503`$ and $`4611`$ filters. The $`2503`$ and $`4611`$ filters were the blocking filters used for the spectroscopy (to limit the extent of the spectra) and are described in Table 1. They are, respectively, a very wide $`V`$ filter and a filter similar to a combination of the $`V`$+$`R`$ filters. The $`V`$ filter is a standard Johnson filter (centered at 5470Å, and FWHM of 880Å) and the $`R`$ filter is similar to the Kron-Cousins $`R`$ but somewhat narrower without the red tail. It is centered at 6500Å, and has a FWHM of 1280Å. From now on we will refer to these filters as $`V_{COP}`$ and $`R_{COP}`$. All fields were imaged for 5 minutes, except for PDCS16 and the two first fields of PDCS38 that were exposed for 15 and 8 minutes, respectively (see Table 3). The images have been photometrically calibrated in the Vega system using several Landolt standard fields (Landolt 1992). Given that most of the fields were observed in only one filter, the photometric calibrations only include a extinction term and a zero point but not a color term. The uncertainties in our photometry were dominated by the fluctuations of the zero points, computed at different airmasses throughout the night, due to imperfect photometric conditions. The internal statistical errors within a field are negligible, except for the faintest objects. We estimate the systematic zero-point error in our measured magnitudes to be less than 0.15 mag for the observations taken in February 1998 (see Table 3) and less than 0.10 mag for the August 1998 observations. Table 3 summarizes our observations. We used the $`V_{COP}`$ exposure for the second field of PDCS62 to complete the $`V_{PDCS}`$ data. This cluster was not completely covered by the PDCS photometry. To transform our magnitudes into a $`V_{PDCS}`$, we have computed the relation between $`V_{PDCS}`$ and our $`V_{COP}`$ for the first field of PDCS62 and applied it to the second field where the $`V_{COP}`$ magnitudes were also available. The best fit obtained was: $`V_{PDCS}V_{COP}=0.28(V_{COP}block_{4611})+0.14`$ We plot finally on Fig. 8 the relations between $`V_{PDCS}`$ and $`I_{PDCS}`$, and between $`V_{PDCS}`$ and $`R_{COP}`$ for all the galaxies in our sample. ### 3.3 Galactic extinction The fields chosen for the PDCS were selected from high-latitude Gun & Oke (1975) survey areas. We avoided regions of high extinction. As expected, the galactic extinction values obtained from the Burstein & Heiles (1982) and Schlegel et al (1998) reddening maps are low. The mean extinction in the $`V_{PDCS}`$ band is 0.027 and lower than this value for the $`I_{PDCS}`$ filter in all the fields (Postman et al 1996). Given the small extinction values and our photometric errors we have chosen not to correct for extinction. ## 4 The CFH spectroscopy ### 4.1 Computing the redshifts We have used both the MIDAS (public ESO reduction package) and IRAF (see e.g. Kurtz & Mink 1998) packages to reduce the 2-Dimensional spectra to 1-Dimensional spectra. The details of the method used can be found in Holden et al. (1999). We computed the redshift from the 1-dimensional spectra from emission lines and from cross-correlation techniques (e.g. Tonry $`\&`$ Davis 1979). If there were more than two emission lines (”only”), we computed the emission line redshift measuring the centroid of the identified lines using gaussian fits and averaging the redshifts. For absorption line dominated spectra, we cross-correlated the spectra with 4 different spectroscopic templates (M31, M32, a 20 Gyear old E/SO Bruzual & Charlot (1993) model and finally a spectrum resulting from the combination of 1959 low-z high quality absorption line spectra: Kurtz & Mink 1998). We used the IRAF/RVSAO package to compute the redshifts. For absorption line dominated galaxies, we produced 4 estimates of the redshift, one from each template. To select a unique value, we proceeded as follows: -1st: eliminated all the redshift estimates lower than -0.015 assuming that even the infalling galaxies of the Virgo cluster have velocities greater than -4500 km s<sup>-1</sup>. This limit is clearly an extreme value. -2nd: eliminated all the redshift estimates with a cross-correlation coefficient lower than 3 (see Kurtz $`\&`$ Mink 1998, Tonry $`\&`$ Davis 1979). -3rd: selected as the true redshift the estimate with the best cross-correlation coefficient if there was a gap of more than 1 with the second best coefficient. -4th: assumed the mean value of the redshifts with the best cross-correlation coefficients if they were in agreement (difference less than 300 km s<sup>-1</sup>). -5th: if neither the 3rd nor 4th conditions were fulfilled, we have simply assumed the value of the redshift with the best cross-correlation coefficient. Using this approach, we obtained 636 redshifts (see Table 2 and 4 for details). We present a sub-sample of 4 spectra in Fig 9. The lower left spectrum is typical of our best signal to noise. It represents a galaxy at z=0.462 with a cross-correlation coefficient of 13.26. The lower right spectrum is a galaxy where we have used emission lines to deduce the redshift (z=0.658), e.g. the \[OII\] line shown on the figure (at $``$6180Å). This spectrum is typical of the worse spectra we used, but the cross-correlation method also was able to detect the CaII H$`\&`$K lines around 6600Å. The upper spectra (left and right) are typical of all our sample. The upper left spectrum is a galaxy at z=0.461 (cross-correlation coefficient of 4.33) and the upper right spectrum is a galaxy at z=0.459 with both absorption line features (cross-correlation coefficient of 3.54) and emission line features (\[OII\]). ### 4.2 Checking the redshifts To check the validity of the assigned redshifts we have ”eye-balled” all the emission line redshifts and also checked a randomly selected sample of absorption line dominated spectra. Our visual inspection confirmed the validity of our computationally derived redshifts. For 19 objects, we also had two separate spectra and, hence, two independent measurements of the redshift. These objects were observed twice due to overlaps in observing runs done at CFH and at the 4 meter Mayall telescope (see Holden et al. 1997). In order to estimate the uncertainties of our redshifts, we have then plotted the percentage difference between the two estimates versus the cross-correlation coefficient $`r`$ (Tonry $`\&`$ Davis 1979). Fig. 10 shows the 18 galaxies with less than 2.5$`\%`$ of difference. The mean error for the redshift estimate is 0.7% of the redshift (or 0.0016 in redshift). The 19<sup>th</sup> galaxy, although having a cross-correlation coefficient of $`r`$=4.2, had a discrepancy of 42% . This is clearly due to one wrong redshift. This galaxy was observed twice at CFHT. The first observation provides a redshift of 0.15364 and a correlation coefficient of 4.2. The second observation yielded a redshift of 0.08958 and a correlation coefficient of 4.2. However, for this observation, the second best estimate (with another template) of the redshift is 0.15673 with a correlation coefficient of 3.71, still acceptable. Using the second value of the redshift, the difference is only 2$`\%`$ for the initially discrepant galaxy. Assuming that these 19 galaxies are representative of our entire sample, we estimate that less than about 5% of our sample has a false redshift assignment. This has a negligible effect on conclusions we draw based on these results. ## 5 Analysis ### 5.1 Final catalogues We identified the objects we measured at CFH (redshift + R<sub>COP</sub> magnitude) with the galaxies in the PDCS survey in order to build catalogues with position (measured at CFH), redshift, R<sub>COP</sub> magnitude and V<sub>PDCS</sub> and I<sub>PDCS</sub> magnitude (Tables 6-15). This is also a way to estimate the uncertainty for the coordinates of the galaxies. We found a mean difference between the PDCS coordinates and the coordinates measured at CFH of 3.5”$`\pm `$2.3”. This is typically the uncertainty for the coordinates we give in table 6-15. We also classified the galaxies in redshift space as members of a structure or as field galaxies. This was a first step. A more detailed classification is discussed in Holden et al. (2000), but we give these results in order to present a complete overview of the data. In order to make this classification, we have searched the velocity distribution of each line of sight for gaps of more than 1000 km s<sup>-1</sup>. If we had more than 5 galaxies between two successive gaps, we have called these galaxies a structure. This is exactly the same method used to define the structures in the ENACS catalog (Katgert et al. 1996). The method does not completely avoid the inclusion of some interloppers, but, to a first approximation, it defines the compact structures (gravitationally bound) in redshift space. We summarize the results in Table 4, Table 6 to 15 and in Figures 11 and 12. ### 5.2 Completeness and spatial representativity We show in Fig. 13 the variation of the completeness level of the spectroscopic catalogue compared to the photometric catalogue. This completeness level $`C`$ is defined as the ratio between the number of galaxies with a measured redshift (galaxies targeted $`and`$ successfully measured) and the total number of galaxies. It is different from the success rate, which is only the ability to deduce the redshift of a target. We see on Fig. 13 that this level is constant around 35$`\%`$ from $`V_{PDCS}`$=16.5 to $`V_{PDCS}`$=21.0. This relatively low level is because we did not have time to put a slit on all the available galaxies. The percentages of the 2 brightest bins of Fig. 13 are based on a low number of galaxies, since, we targeted more faint galaxies than bright galaxies explicitely to try to keep constant the completeness level $`C`$. This constant sampling is important for studying the galaxy distribution along the line of sights because it prevents us from severe redshift selection effects. For the faintest galaxies, the completeness level drops down to 6$`\%`$ for $`V_{PDCS}`$=22.5. These percentages were computed using all the lines of sight put together except PDCS61 for which the exposure time was very short. These percentages do not change considerably from pointing to pointing (except for PDCS61). For several types of analysis, it may be useful to give an analytical expression of the variation of the completeness level $`C`$. For the magnitudes brighter than $`V_{PDCS}`$=20.5, $`C`$=35.5$`\%`$. For the fainter magnitudes, assuming a power law model, the best fit is: $`C=10^{0.44(V_{PDCS}24.07)}\%`$ To test for selection effects in the spatial distribution of the galaxies for which we measured redshift, we compared the spatial distribution of our redshift measured sample to that of all the PDCS galaxies. For this comparison, we have used a bidimensional Kolmogorov-Smirnov test as a function of the V limiting magnitude to determine the variation of this representativity level as a function of the photometric depth of the sample. A value of the representativity level given by the Kolmogorov-Smirnov test close to 100$`\%`$ means that the two distributions on the sky are very similar: the galaxies with a measured redshift are a statistically representative sub-sample in term of spatial distribution. A value lower than 90$`\%`$ means that this sub-sample is statistically different at the level of 10$`\%`$. We see in Tab. 5 that, except for PDCS30-45, the two spatial distributions are indistinguishable for the magnitudes brighter than 21. Note that the case of PDCS30-45 is not easely explained (see Tab. 5). ## 6 Summary We have presented and given the data gathered in the COP survey. The spectroscopic and photometric observations were performed with the MOS/STIS2 instrument during 6 nights at the CFH telescope with the grism O300 and 2 blocking filters to enhance the multiplex gain of MOS. We have used a method based on the MST theory to optimize the number of slits per mask. This allowed us to measure 636 redshifts for 10 PDCS lines of sight. These lines of sight were selected to hold PDCS candidate clusters, with significant peaks in the galaxy density distribution. The success rate (percentage of targeted galaxies with a successfully measured redshift) was close to 70$`\%`$ for the primary targets (typically brighter than V<sub>PDCS</sub>=22.). The completeness level (percentage of all galaxies with a measured redshift) was about 35$`\%`$ down to V<sub>PDCS</sub>=20.5. The galaxies with a redshift were proved to be a spatially representative sub-sample down to V<sub>PDCS</sub>=20.5 (no significant spatial selection effects). Finally, the percentage of false redshifts was about 5$`\%`$, based on 19 galaxies observed twice. A comparison of the photometry from the PDCS (Postman et al. 1996) catalogs and from the new images we have obtained at the CFH telescope shows that the different magnitude systems can be cross-calibrated. This confirmation is important for the reliability of future works based on the multi-color photometry of COP. After identification between the PDCS catalogues and our new images, we built catalogues with redshift, coordinates and V<sub>PDCS</sub>, I<sub>PDCS</sub> and R<sub>COP</sub> magnitude (Tab. 6-15). We have classified the galaxies along the lines of sight into field and structure galaxies using a gap technique (Katgert et al. 1996). In total we have observed 18 significant structures along the 10 lines of sight (Tab. 4). As noted in the introduction, the interpretation of the results is given elsewhere (e.g. Holden et al. 2000). CA thanks the staff of the Dearborn Observatory for their hospitality during his postdoctoral fellowship. The authors thanks the CFHT TAC for support. BH would like to acknowledge support from the following: NSF AST-9256606, NASA grant NAG5-3202, NASA GO-06838.01-95A, and the Center for Astrophysical Research in Antarctica, a National Science Foundation Science and Technology Center.
warning/0003/cond-mat0003021.html
ar5iv
text
# Coulomb Charging Effects in an Open Quantum Dot ## I Introduction Charge quantization plays a central role in electron transport through lateral quantum dots weakly coupled to leads. It had been commonly believed, argued by justification of validity of Coulomb blockade theory and various experimental results, that Coulomb blockade oscillations of the conductance weaken and gradually vanish as the transparency of the barriers increases up to conductance quantum $`2e^2/h`$. However, the possibility of Coulomb charging in open quantum dots is now intensively investigated both theoretically and experimentally. In strong magnetic fields continuous Coulomb oscillations superimposed upon large-period oscillations have been detected with the conductance $`G`$ ranging up to $`G=3e^2/h`$. Only recently experimental evidence of single-electron charging of an open quantum dot has been obtained in zero magnetic field. In Ref. the observation of pairs of resonant peaks in each transition region between the quantized conductance plateaus have been reported for the conductance of 1D channel changing up to $`G=7(2e^2/h)`$. In this case the potential of an impurity formed a small quantum dot in the channel, and the resonant peaks were ascribed to Coulomb charging of the dot. Detailed studies of this effect were, however, not possible because the transparency of the barriers could not be precisely controlled. Recently novel type quantum dots with overlaying finger gates were fabricated. Surprisingly, continuous and periodic oscillations superimposed upon ballistic conductance steps were observed when the conductance through the dot changed within a wide range $`0<G<6e^2/h`$. A smooth transition of the oscillations from $`G>2e^2/h`$ to $`G<2e^2/h`$ with decreasing barrier transparency leads to the conclusion that the oscillations are due to single-electron charging of the quantum dot. However, none of the existent theories can explain the manifestation of single-electron oscillations over such a wide range of the conductance. In this paper we analyze the results of conductance measurements of this novel type of quantum dot at zero magnetic field and discuss some new observations. Additionally, we report realistic modelling of the electrostatics and electron transport in the quantum dot. By calculating the capacitances of the quantum dot with respect to the gates we confirm the single-electron origin of the weak conductance oscillations. Results from the modelling of the electron transport show that mixing of the 1D-subbands is almost absent and that the large-scale resonant features in the background conductance are due to Fabry-Pérot interference. Thus in these devices single-electron charging and coherent electron transmission at $`G>2e^2/h`$ coexist. In the present experiment the quantum dot was defined by two side gates, which deplete electrons within the channel, and three narrow overlaying finger gates. Outermost finger gates introduce the entrance and exit barriers to the dot, and the central finger gate stabilizes the depth of the potential inside the dot. The impurity scattering in the device, fabricated on an ultra-high-quality high electron mobility transistor (HEMT), is negligible. Our calculations demonstrate the unique versatility of this dot geometry with adjustable voltages on the side gates and three finger gates. We show that in some voltage regimes the electrostatic potential in the plane of the two-dimensional electron gas (2DEG) is separable as $`U(x,y)=U_1(x)+U_2(y)`$ and thus the device exhibits simple one-dimensional behaviour. In more standard quantum dots where the constrictions are defined by two pairs of split gates, Coulomb oscillations in zero magnetic field are only observable at $`G<2e^2/h`$. In order to determine the difference between this new, versatile device and the more standard quantum dot we compare the calculated electrostatics and transport in the two different types. Our results show that inter-1D subband scattering is suppressed in the new type of open quantum dot, owing to the special design, whereas in the more standard quantum dots the intersubband mixing is considerably enhanced once the transmission via the first subband is opened. We argue that quasi-1D transport through the quantum dot and high sensitivity of the barrier transparency in the constrictions to the variations of the Fermi level in the dot makes it possible to observe the effects of Coulomb charging at $`G>2e^2/h`$. The paper is organized as follows. In Sec. II the quantum dot device and conductance measurements are described. The behaviour of large scale features and frequent oscillations of the conductance with gate voltages and temperature are analyzed in details. Numerical results are reported in Sec. III. First we discuss the electrostatics of the device and determine the capacitance of the dot with respect to the contacts, and finger and split gates. Then the calculated two-dimensional potential profile was used for modelling multiple mode electron transmission through the quantum dot. In Sec. IV we give qualitative account for the observed single-electron conductance oscillations. ## II Experiment ### A Structure characterization <br>and main effect The two-layered Schottky gate pattern shown in Fig. 1 was defined by electron beam lithography on the surface of a high-mobility GaAs/Al<sub>0.33</sub>Ga<sub>0.67</sub>As heterostructure T258, 157 nm above a 2DEG. There is a 30-nm-thick layer of polymethylmethacrylate (PMMA) which has been highly dosed by an electron beam, to act as a dielectric between the split gate (SG) and three gate fingers (F1, F2, and F3) so that all gates can be independently controlled. After brief illumination by a red light emitting diode, the carrier concentration of the 2DEG was $`1.6\times 10^{15}`$ m<sup>-2</sup> with a mobility of $`250`$ m<sup>2</sup>/V s. The corresponding transport mean free path is $`16.5\mu `$m, much longer than the effective 1D channel length. Experiments were performed in a dilution refrigerator at $`T=50`$ mK and the two-terminal conductance $`G=dI/dV`$ was measured using an ac excitation voltage of $`10`$ $`\mu `$V with standard phase-sensitive techniques. In all cases, a zero-split-gate-voltage series resistance ($`900\mathrm{\Omega }`$) is subtracted. Two samples, at five different cooldowns, show similar characteristics, and measurements taken from one of these are presented in this paper. Trace 1 in Fig. 2 shows the conductance measurements $`G(V_{\mathrm{SG}})`$ as a function of split-gate voltage $`V_{\mathrm{SG}}`$ when all finger gate voltages $`V_{\mathrm{F1}}`$, $`V_{\mathrm{F2}}`$, and $`V_{\mathrm{F3}}`$ are zero. Conductance plateaus at multiples of $`2e^2/h`$ are pronounced (with no resonant feature superimposed on top) as expected for a clean 1D channel. When the channel is defined at $`V_{\mathrm{SG}}=1.132`$ V, six quantized conductance steps are observed when each one of the finger gates is swept while the others are grounded to the 2DEG as shown in traces 2–4 (Fig. 2). These experimental results demonstrate that a clean 1D channel is obtained in which impurity scattering is negligible. A lateral quantum dot was defined by applying voltages on SG, F1, and F3, while keeping F2 grounded to the 2DEG. Resonant features are observed only when large negative voltages are applied to both F1 and F3. With some depletion voltage $`V_{\mathrm{F1}}V_{\mathrm{F3}}2`$ V at low-temperature, almost periodic and continuous oscillations of conductance $`G(V_{\mathrm{SG}})`$ over a wide range $`0<G<6e^2/h`$ are observed. Typical traces of the conductance $`G`$ and the distance between adjacent peaks (“period”) $`\delta V_{\mathrm{SG}}`$ as functions of gate voltage $`V_{\mathrm{SG}}`$ are shown in Fig. 3. The period and shape of the oscillations remain approximately the same within the wide range where $`V_{\mathrm{SG}}`$ is varied, though the background conductance changes considerably. While the oscillations at $`G<2e^2/h`$ can be ascribed to Coulomb charging effects, the oscillations for $`G>2e^2/h`$ are unexpected. The observed oscillations are essentially different from the single-electron effects in older type lateral quantum dots in which single electron tunneling peaks increase in height and decrease in width as the conductance decreases, and the period of the oscillations is well defined. The latter behaviour corresponds to the orthodox theory of Coulomb blockade. In contrast to the majority of the other papers where low-temperature single-electron effects have been studied, the trace in Fig. 3 has neither equally spaced narrow peaks nor regions of strongly suppressed conductance (Coulomb blockade) inbetween the peaks. Instead, all the oscillations in Fig. 3 are smoothed, have small amplitude $`0.2e^2/h`$, are approximately the same width, and the peak spacing fluctuates by several tens of percent. Figure 4 shows that similar oscillations are observed when the central finger gate voltage $`V_{\mathrm{F2}}`$ is varied, with the side gate voltage fixed. At the top of Fig. 4 the conductance $`G`$ is shown along with its running average $`G_{\mathrm{RA}}`$ (the *background*). The oscillations without the background, $`GG_{\mathrm{RA}}`$, are shown at the bottom and their Fourier spectrum is given in the inset. Noticeably, in the region $`V_{\mathrm{F2}}<0.6`$ V, where $`G<e^2/h`$, the peaks should be strictly equidistant according to Coulomb blockade theory, additional beats are evident. The observed oscillations overlay the wide maxima or steps of the conductance which appear periodically with the central finger gate voltage $`V_{\mathrm{F2}}`$ varied at fixed $`V_{\mathrm{SG}}`$ (Fig. 4) and are most likely associated with electron wave interference on the system of two barriers in the constrictions. With changing the depth of the potential well in the quantum dot the resonances move through the Fermi level one by one. Previous studies have reported the observation of Coulomb oscillations superimposed on almost periodic conductance peaks at $`G<e^2/h`$ and were interpreted as Fabry-Pérot resonances due to coherent electron tunneling through the quantum dot. In our case the frequent small-amplitude oscillations penetrate to the region $`G>2e^2/h`$, where the transport is traditionally considered coherent and Coulomb charging effects are not usually observed. Generally for $`G>2e^2/h`$, it is expected that the presence of a fully transmitted 1D channel might cause mode mixing between 1D channels in the quantum dot which should smear out charging effects. However, due to the special design and high quality of this device it is likely that there is little 1D mode mixing in the chosen range of gate voltages such that the level broadening for Coulomb oscillations is similar for both cases when $`G<2e^2/h`$ and $`G>2e^2/h`$. ### B Temperature evolution Figure 5 shows how the features in the background (a) and the oscillations (b) of the conductance of the quantum dot develop with decreasing temperature from $`1`$ K to $`50`$ mK. Consider the behaviour of the background first \[see Fig. 5(a)\]. At 1 K there is no conductance quantization. Only the wide shoulder at $`V_{\mathrm{SG}}<0.65`$ V marks out the tunneling regime of the first subband $`G<e^2/h`$. Around $`T=0.2K`$ two plateaus at $`0.8(2e^2/h)`$ and $`1.8(2e^2/h)`$ appear. With lowering the temperature down to 50 mK resonant features develop: plateaus transform to peaks and shoulders emerge at $`0.7(2e^2/h)`$ and $`1.2(2e^2/h)`$. Thus, in contrast to the well-quantized ballistic conductance plateaus shown in Fig. 2, applying voltages to F1 and F3 results in conductance steps that are not as flat or well quantized. In figure 5(b), at 1 K only a group of $`10`$ weak conductance oscillations within the range $`0.7\text{V}<V_{\mathrm{SG}}<0.65`$ V are discernible. These oscillations are located at the bottom slope of the first subband $`G(V_{\mathrm{SG}})`$ and relate to the tunneling regime $`G<e^2/h`$, so they can be ascribed, by analogy with Coulomb blockade peaks, to Coulomb charging effects. For $`T=0.5\text{}0.4`$ K a group of $`6`$ oscillations appear in the range $`0.65\text{V}<V_{\mathrm{SG}}<0.625`$ V which ascend up to $`G2e^2/h`$. Then at $`T=0.26\text{}0.2`$ K oscillations at higher conductance $`2e^2/h<G<4e^2/h`$ show up for $`0.625\text{V}<V_{\mathrm{SG}}<0.53`$ V. And finally, at $`T=0.15\text{}0.11`$ K oscillations become visible when the conductance is between the second and third quantum $`4e^2/h<G<6e^2/h`$. This division into groups of oscillations is traced down to $`T=0.05`$ K. Within the groups, additional modulation of the amplitude of oscillations is pronounced and correlates with the resonant features of the background conductance $`G_{\mathrm{RA}}`$. Thus, each group can be characterized by the temperature at which its oscillations become visible. The fact that the frequent oscillations appear at $`T=1\text{}0.5`$ K before the wider resonant features do at $`T<0.2`$ K precludes their unified interpetation by transmission resonances at coincidences of quasidiscrete levels of the dot with Fermi level. Indeed, the energy scale, i. e. critical $`k_BT`$ of the frequent oscillations exceeds that of the wider features, while the resonances at the single-particle levels would have smaller energy spacing than the width of Fabry-Pérot resonances. The level spacing estimated from Aharonov-Bohm oscillations following the method described in Ref. gives $`\mathrm{\Delta }E12\mu `$eV, comparable to the thermal smearing at 150 mK. On the other hand, estimating the charging energy from the critical temperature $`T1`$ K at which the oscillations are still observed results in $`e^2/2C_\mathrm{\Sigma }0.2`$ meV. Thus the observed oscillations are not due to resonances on single-particle levels of the dot and can be described in terms of the Coulomb charging picture where the 0D quantum confinement energy is much smaller than the Coulomb charging energy. The most likely explanation for the frequent oscillations showing up in the region $`G>2e^2/h`$ with lowering temperature is the decrease of the decay probability of the localized states in the dot via fully transmitted 1D-subbands. Increasing the temperature enhances the mixing between transmitted and closed 1D-subbands such that oscillations suppressed at $`G>4e^2/h`$ (two 1D-subbands are fully transmitted), thereafter become suppressed at $`G>2e^2/h`$ (1D-subband), and eventually at $`G<e^2/h`$, where only tunneling decay of the localized states is possible. ### C Gate voltage dependences of oscillations Figure 6 demonstrates how the background conductance $`G_{\mathrm{RA}}(V_{\mathrm{SG}})`$ (a) and the oscillations less the background (b) change with incremental voltage steps on the outermost finger gates $`V_{\mathrm{F1},3}`$ at $`T=50`$ mK. One can see that the oscillations of $`G(V_{\mathrm{SG}})`$ gradually evolve with similar shape and periodicity from the region $`G>2e^2/h`$ to the region $`G<2e^2/h`$, indicating their common physical origin. Since for $`G<2e^2/h`$ Coulomb blockade theory holds and the oscillations are due to Coulomb charging, the observed evolution of the oscillations to $`G>2e^2/h`$ serves as an experimental confirmation of the single-electron nature of all the oscillations. From the observed Aharonov-Bohm type oscillations the dot area was determined to be $`A=2.81\times 10^{13}`$ m<sup>2</sup> and the number of electrons was $`n=126`$ for $`V_{\mathrm{SG}}=0.5`$ V, $`V_{\mathrm{F1}}=1.941`$ V, and $`V_{\mathrm{F3}}=1.776`$ V. Provided that every conductance oscillation corresponds to a change of the dot charge by $`e`$, one can determine (from the total number of oscillations $`\mathrm{\Delta }n50`$) that there are still $`70`$ electrons within the dot at pinch-off $`V_{\mathrm{SG}}=0.65`$ V. Note that the modulation of the oscillation amplitude in Fig. 6(b) allows us to trace the movement of the oscillations for different $`\left(V_{\mathrm{SG}},V_\mathrm{F}=(V_{\mathrm{F1}}+V_{\mathrm{F3}})/2\right)`$ along almost parallel lines which corresponds to the conservation of the dot charge $`Q=ne`$. Two such lines are shown dashed in Fig. 6(b); where both horizontal and vertical separation between them obeys the same condition $`\delta Q=e`$. Wide stripes of amplitude modulation markedly divide the oscillations into groups in agreement with Fig. 5(b), and their slope reproduces that of the conductance threshold. Also, the groups of oscillations correlate with the locations of the resonant features of the background $`G_{\mathrm{RA}}`$ in Fig. 6(a). The background conductance $`G_{\mathrm{RA}}(V_{\mathrm{SG}})`$ \[see Fig. 6(a)\] contains both steps and bumps which move with the conductance threshold as the finger gate voltage raises the barriers in the constrictions; these features are smeared out and completely disappear with decreasing transparency of the barriers. There are several reasons for such behaviour. Firstly, the tops of the barriers in the constrictions approach the Fermi level with a large negative voltage on the finger gates. Any small, but inevitable, asymmetry between the two constrictions on the transport properties will be enhanced, causing the height of the steps and resonant peaks of the conductance to reduce. Secondly, smearing of the features that occurs at lower negative voltages $`V_{\mathrm{SG}}`$, when $`V_{\mathrm{F1},3}`$ increases, is favoured by a widening of the constrictions over this voltage range thereby reducing the intersubband spacing there. And lastly, any decrease of the transparency of the barriers increases the electron dwell time in the dot and the role of decoherence, such that the constrictions start acting independently. The quantised coductance steps for a single constrictions smear out in this voltage regime as shown by both measurements \[see Ref. , Fig. 2(a)\] and modelling. The condition $`\delta Q=e`$ can be used to find the capacitances between the dot and the gates of the sample. For this purpose the periods of conductance oscillations versus gate voltages were measured, $`\mathrm{\Delta }V_{\mathrm{F1}}=23.8`$ mV, $`\mathrm{\Delta }V_{\mathrm{F2}}=8.7`$ mV, $`\mathrm{\Delta }V_{\mathrm{F3}}=25.9`$ mV, and $`\mathrm{\Delta }V_{\mathrm{SG}}=3.6`$ mV. According to this the total gate-dot capacitance $`C_g`$ is estimated to be $`7.6\times 10^{17}`$ F. ## III Numerical results ### A Electrostatics In order to check the correspondence between the observed period of oscillations and the change of the dot charge by one electron, and to obtain an estimate of the charging energy, we calculated the capacitance of the dot with respect to the contacts, fingers, and split gates. The electrostatic potential profile in the device was determined by solution of the 3D Poisson equation with a local 2DEG density given by the 2D Thomas-Fermi approximation assuming a boundary condition of frozen charge at the surface states and impurities. It was checked that fluctuation potential in this structure due to ionised dopants is absent due to the wide AlGaAs spacer ($`100`$ nm). The conformity of this fairly simple model to the experiment was checked by calculation of the pinch-off voltages. The calculated Fermi level $`E_F=5`$ meV in the 2DEG reservoirs corresponds to the measured carrier density $`n=1.6\times 10^{11}`$ cm<sup>-2</sup>. In the calculations the same voltage $`V_\mathrm{F}`$ was applied to the outermost finger gates, and the central finger gate was set at zero voltage. At $`V_\mathrm{F}=0`$ the channel pinches off when $`V_{\mathrm{SG}}=1.8`$ V (the same as in the experiment). When $`V_{\mathrm{SG}}=0.7`$ V the finger gates raise the potential barriers in the constrictions above the Fermi level at $`V_\mathrm{F}=1.4`$ V (experimentally the split-gate pinches off at $`V_{\mathrm{SG}}=0.7`$ V when $`V_{\mathrm{F1}}=1.9`$ V and $`V_{\mathrm{F3}}=1.7`$ V). We ascribe this small difference between the calculated ($`V_\mathrm{F}`$) and experimental ($`V_{\mathrm{F1}},V_{\mathrm{F3}}`$) values to the fact that we do not take into account the capacitances of the finger gates with respect to the shield of the structure (we also neglect electric field lines going above the PMMA layer). We calculated the potential profile, charge distribution and the total charge of the dot, as well as the capacitances in the range $`V_{\mathrm{SG}}=0.75`$ to $`0.5`$ V and $`V_\mathrm{F}=1.3`$ to $`1.4`$ V which closely agrees with the range of experimental gate voltages specified for the traces in Figs. 35. These results are shown in Figure 7 which shows maps of the charge density in the quantum dot for closed and open states. With lowering $`V_{\mathrm{SG}}`$ the dot stretches along $`y`$ axis and becomes rectangular. Transverse cross sections of the electrostatic potential in the 2DEG are shown in Fig. 8(a,b) for two different $`x`$ coordinates along the channel: in the center of the dot at $`x=0`$, and directly beneath the finger gates at $`x=270`$ nm. By changing the voltage $`V_{\mathrm{SG}}`$ the dot transforms from a closed state (a) to an open state (b), with a corresponding change in the width of both the dot and the constriction. The voltage on the finger gates control both the height of the barriers and the width of the constrictions (Fig. 9) with little change in the depth of the dot. At large finger gate voltages $`V_\mathrm{F}=1.3`$ to $`1.4`$ V and low side-gate voltages $`V_{\mathrm{SG}}0.5`$ V the transverse potential profile of the constriction resembles a rectangular well (Fig. 9). With the central finger gate kept at zero voltage, the width and depth of the quantum dot were found to depend on $`V_{\mathrm{SG}}`$ only. It is interesting to note that the presence of zero-biased F2 makes the dot 0.5 meV deeper in energy and stabilizes its depth at $`3`$ meV \[see Fig. 8(a,b)\]. On the other hand, if the voltage on the outermost finger gates is fixed and the central finger gate voltage is varied, it mainly changes the depth of the potential in the quantum dot (Fig. 10). The calculations show that the number of electrons in the dot changes from 80 to 140 as the side gate voltage changes from $`V_{\mathrm{SG}}=0.7`$ V to $`0.5`$ V (with fixed $`V_{\mathrm{F1}}=V_{\mathrm{F3}}=1.3`$ V). This change in the number of electrons corresponds to the number of oscillations observed in Fig. 3. Calculated capacitances of the dot to the gates are also close to the experimentally estimated ones and lay within the measured period variation (Table I). In calculations the capacitances demonstrated the same systematic drift with $`V_{\mathrm{SG}}`$ and $`V_\mathrm{F}`$ as observed in Fig. 3(b). Thus, the conclusion that each oscillation of the conductance reflects the change of the dot charge by one electron is confirmed. By introducing a small Fermi level difference between the dot and the 2DEG reservoirs, we calculated the capacitance of the dot with respect to both contacts as $`C_r=340\text{}370`$ aF for an almost closed quantum dot. The capacitance is doubled when three 1D-subbands become transmitted. Thus, this capacitance is almost an order of magnitude higher than that to the gates and cannot be neglected, so the charging energy is $`e^2/2C=0.1\text{}0.2`$ meV, where $`C=C_r+e/\mathrm{\Delta }V_{\mathrm{SG}}+e/\mathrm{\Delta }V_\mathrm{F}+e/\mathrm{\Delta }V_{\mathrm{F2}}`$, comparable to the thermal broadening at $`T1\text{}2`$ K. As Fig. 5 shows, near the pinch-off the conductance oscillations persist up to 1 K, in accordance with the conventional theory of Coulomb blockade. The decrease of the charging energy to 0.1 meV at $`G6e^2/h`$, as found in the calculations of the electrostatics, should lower the limiting temperature for observing the oscillations in this range to $`0.5`$ K. Figure 5 shows that in reality the measured temperature is still 3 times smaller. This strong reduction could be caused by an enhanced decay of the localized states via two fully transmitted subbands and an increase in the intersubband mixing. ### B Comparing quantum dots of different types To understand the difference between the dot under study \[Fig. 11(a)\] and a more standard quantum dot (where the constrictions are induced by two pairs of split gates and Coulomb oscillations are observed only at $`G<e^2/h`$) calculations of the electrostatics were also carried out for the case in which the outermost 160 nm wide finger gates were separated by a 260 nm gap \[Fig. 11(b)\]. We will denote those devices as A and B, respectively. Except for the finger gates, all the parameters of devices A and B are the same. Calculated capacitances of the quantum dot in closed and open states for cases A and B are similar. The essential differences between the electrostatic potentials in the plane of the 2DEG only appear in the constrictions. In device B, the barriers in the constrictions $`x=x_c`$ are lower, and the transverse cross section of potential there resembles a deep and narrow parabola $`U(x_c,y)=U_c+m\omega _c^2y^2/2`$ with energy quantum $`\mathrm{}\omega _c=0.6\text{}0.8`$ meV \[Fig. 11(d)\]. The quantum in the centre of the channel $`x_d=0`$ (the quantum dot) is 2–3 times smaller: $`\mathrm{}\omega _d=0.2\text{}0.3`$ meV. In device A the transverse potential in the constriction resembles a cut parabola \[Fig. 11(c)\], so the lowest 1D subbands are denser near the bottom, like that in a rectangular potential well. When the quantum dot is open for transmission via the first subband, the 1D subband spacing in device A is almost equal both inside the dot and constrictions: $`E_{n+1}E_n=0.2\text{}0.3`$ meV. The energy levels of transverse quantization $`E_n(x)`$ were determined from a solution of the Schrödinger equation for the calculated electrostatic potential $`U(x,y)`$ by a tight-binding method. To impose zero boundary conditions for transverse motion, infinite walls were put at 600 nm from the axis of the channel. The picture of 1D-subbands shows how the subband spacing changes along the channel axis and how many subbands are open for transmission through the quantum dot at a given Fermi level. Figures 11(e,f) show the positions of three lowest 1D-subbands $`E_n(x)`$ for devices A and B. The Fermi level is shown by a dotted line and corresponds to zero energy. In case A, the subband spacing is almost independent of $`x`$. This means that the transverse cross sections of potential in the dot and in the constrictions have the shape of the same parabola, in other words $`U(x,y)=U(x)+m\omega _c^2y^2/2`$. Then the variables $`x`$ and $`y`$ in the Schrödinger equation are separated, and the motion along $`x`$ and $`y`$ directions is described by separate equations, with no mixing between different 1D subbands. Thus the transmission problem reduces to one-dimensional one. Contrarily, in device B, where there is a gap between the finger gates, 1D subbands are not parallel and the intersubband spacing changes by 2–3 times along the channel \[Fig. 11(f)\]. Thus the potential has such a shape that the variables in the Schrödinger equation cannot be separated, the mixing between 1D-subbands is strong and the motion is essentially two-dimensional. Electron transmission can only be considered one-dimensional when the first subband is opening and the transmission coefficient $`T<1`$. These assumptions about one-dimensional transmission in device A and two-dimensional transmission in device B are supported by numerical calculations of multiple-mode transmission, as described in the next sub-section. ### C Electron transmission through quantum dots Two-dimensional transmission was calculated on the same grid in variables $`(x,y)`$ as the Poisson equation was solved for $`U(x,y)`$. Along the channel axis $`x`$, energy levels $`E_n`$ in each transverse cross section and transfer matrix elements between adjacent cross sections were determined and then the multiple-mode transmission problem was solved by means of scattering $`S`$-matrices. The conductance relates to the total transmission coefficient according to the Landauer formula: $$G=\frac{2e^2}{h}T,T=\underset{n}{}T_n,T_n=\underset{k}{}|T_{nk}|^2.$$ The transmission was calculated for quantum dots and single constrictions (half the quantum dot). In Fig. 12(a) plots of the Fermi energy dependence of the total transmission coefficient and its modal contributions are shown for device A ($`V_{\mathrm{SG}}=0.49`$ V, $`V_\mathrm{F}=1.4`$ V). The dashed lines show the transmission through single constrictions. When the first mode is 50% transmitted, the second mode has already reached 30% and so on. For small values of the transverse quantum $`\mathrm{}\omega =E_2E_1=0.2`$ meV, conductance quantization is smeared out on a single constriction, though it can occur for resonant transmission through two barriers in series. Transmissions in the quantum dot are shown by the solid lines in Fig. 12. For device A the transmission curves for the first to third subbands resemble each other but with an offset in energy by the transverse quantum. Similar behaviour is observed in the split gate voltage dependence of the transmission in Fig. 13 which models the experimental situation shown in Fig. 3. It is important that only a few resonant features are present in $`T_n(E)`$ and $`T_n(V_{\mathrm{SG}})`$ — those are Fabry-Pérot resonances in the system of two barriers. The narrowest ones of the resonances are marked with triangles and refer to the tunneling regime of the corresponding subbands; they are smeared out in measurements and not visible in Fig. 3 since in this regime the transport is sequential rather than coherent. Contrarily, the wide resonances refer to above-barrier coherent transmission (marked with asterisks) and give rise to every next step of conductance quantization in Figs. 35. It seems that the conductance steps in Fig. 3 are not the property of a single barrier, but the property of the pair of barriers \[see Figs. 13 and 12(a)\]. The difference in the height of the barriers of $`0.1\text{}0.2`$ meV (weak asymmetry of the structure) causes no subband mixing but reduces the conductance steps and shifts the resonances (e. g. dotted curve in Fig. 12). This asymmetry can explain the observed transformation of background conductance $`G_{\mathrm{RA}}(V_{\mathrm{SG}})`$ in Fig. 6(a). The total transmission coefficient and modal distribution for device B are shown in Fig. 12(b) ($`V_{\mathrm{SG}}=0.5`$ V, $`V_\mathrm{F}=1.6`$ V). Dotted curves show transmission with pronounced steps for a single constriction. Because of the large subband spacing the tunneling in closed subbands is negligible. In transmission through the dot, however, the transport may go via the higher subbands due to mixing with lower open subbands. For instance, nonzero transmission via the third subband occurs due to coupling to the first subband even if the second subband is not yet transmitted. When $`2e^2/h<G<6e^2/h`$, the transport involves more than five modes and higher modes contribute much more to the conductance than those in device A. The intersubband mixing shows up in the $`T(E_F)`$ dependence as sharp Fano resonances due to electron scattering from the levels of the dot \[Fig. 12(b)\]. The dependence $`T(V_{\mathrm{SG}})`$ is similar to $`T(E_F)`$: transmission is one-dimensional at $`T<1`$ and already multimodal at $`T>1`$. It should be noted that while 1D-subbands in device A become absolutely transparent ($`T_n=1`$) with increasing energy or $`V_{\mathrm{SG}}`$, the transparency of open subbands in device B changes resonantly from 0 to 80–90% due to intersubband mixing. This can explain why charging effects are smeared out at $`T>1`$ in more standard quantum dots. Figure 14 shows the modelled dependence of the conductance on the central finger gate voltage $`G(V_{\mathrm{F2}})`$. The corresponding deformation of the potential in the dot and constrictions was shown in Fig. 10. The depth of the potential in the dot decreases, and the resonances due to 1D interference on the two barriers cross the Fermi level one by one. The calculated coherent transmission is shown in Fig. 14 by the solid curve from which the dashed curve without sharp peaks is obtained by smoothing. Five wide Fabry-Pérot resonances are clearly seen in the figure which are also present on the experimental curves (the background in Fig. 4). The fact that the number of frequent oscillations in experimental curves (Figs. 3, 4) differs drastically from the number of resonant features in the calculated transmission coefficients (Figs. 13, 14) demonstrates that the observed frequent oscillations are not due to interference effects of coherent electron transmission through quasi-discrete states of the quantum dot. In the coherent regime an electron does not scatter on most levels in the absence of mode mixing. The suppression of mode mixing is a consequence of the geometry of the dot and the corresponding selected range of voltages at finger and side gates. Large negative voltage at overlaying finger gates flattens the potential across the channel so that the separation between the lowest 1D-subbands in the constrictions becomes as small as that in the dot \[Fig. 11(e)\]. Calculation of the transmission coefficients show that mode mixing is strengthened when the conductance rises to $`G6e^2/h`$. This explains why the amplitude of the measured conductance oscillations and the temperature at which the oscillations vanish are reduced in this voltage range. ## IV Qualitative explanation of oscillations Based on the modelling of the electrostatics and coherent transmission we suggest the following scenario of Coulomb charging in an open quantum dot. There are three important features that the new type of dot possesses: 1) coupling between localized and transmitted subbands is suppressed; 2) coherent transmission and sequential tunneling coexist; 3) the charging energy and 1D-subband spacing in the constrictions are commensurate. We presume that in the absence of intersubband mixing the transport in the low-transparency subbands is due to sequential tunneling with single-electron charging of the dot, resulting in the conductance oscillations (Figs. 36). The open subbands transmit the electrons coherently and provide a parallel background current with Fabry-Pérot resonances and quantization steps in the conductance. With each new step the single-electron charging in the opening subband ceases but it still takes place in a higher (low-transparency) subband, since the intersubband transition probability in the dot is very small at low temperature. Apparently, when the transmissions via different 1D-subbands are independent, Coulomb charging effects are manifested in a similar manner for each opening subband. That the experimentally observed conductance oscillations are continuous is due to the small spacing of transverse quantization levels in the constrictions, leading to stronger tunneling via closed subbands compared to more standard quantum dot devices. The fact that the spacing of 1D-subbands $`\mathrm{}\omega `$ in the constrictions is approximately equal to the charging energy $`e^2/2C=0.1\text{}0.2`$ meV can account for the uniformity and smoothness of the observed oscillations. Indeed, electron localization and tunneling makes the charge on the dot follow the quantization, i. e. the dependence of the dot charge on gate voltage departs from linearity proportionality towards the step function \[Fig. 15(a)\]. From electrostatics it follows that the deviation of the dot charge $`Q`$ from the value $`C_gV_g`$ produces a voltage difference $`V_b`$ between the dot and reservoirs. Precise charge quantization, if it were in the Coulomb blockade regime, would lead to sawtooth modulation of $`V_b`$ between $`V_b=e/2C`$ and $`V_b=e/2C`$ as the value $`C_gV_g`$ changes by $`e`$ \[Fig. 15(b)\]. However, because of the condition $`e^2/2C\mathrm{}\omega `$ the transparency of the barriers in the constrictions varies strongly with changing $`V_b`$ and thus causes a periodic change of that part $`q`$ of the dot charge which is associated with the population of the delocalized states. As a result of continuous change of $`q`$ ($`|q|<e/2`$) the steps and sawteeth of the gate voltage dependences $`Q(V_g)`$ and $`V_b(V_g)`$ are smoothed. Nevertheless, if the decay rate from the localized states to the transmitted ones is low these features survive even if the transport is coherent and fully transmitted subbands are present. One may imagine that charge $`q`$ plays the same role as the polarization charge of the Coulomb island plays in the conventional theory of Coulomb blockade. In this theory, when parameter $`q`$ is kept constant, the charge becomes strictly quantized at zero temperature. However, temporal fluctuations of $`q`$ widen the sharp features in $`Q(V_g)`$ and $`V_b(V_g)`$. In our case the situation is similar. We have numerically found that the transparency of the quantum dot changes by $`0.3(e^2/h)`$ when $`V_b`$ is varied by only $`0.1e/C`$ (the bottom of the dot is raised by $`0.1e^2/C`$). Thus, the periodic change of the embedded voltage $`V_b`$ with gate voltage results in single electron conductance oscillations in the coherent current \[Fig. 15(c)\]. Because the charging energy and the subband spacing in constrictions are approximately equal it follows that there is no principal difference between regimes $`G>2e^2/h`$ and $`G<2e^2/h`$. In reality, besides coherent transmission there is sequential tunneling that leads to spontaneous switching between adjacent charge states of the dot. Thus for $`G<2e^2/h`$ significant time is spent in a charge state with highly transparent potential barriers and similarly, for $`2e^2/h<G<4e^2/h`$ the dot often happens to be in a charge state with a low transparency of the barriers. Consequently the observed charge oscillations of the conductance appear uniformly smoothed. In addition, the amplitude and the period of such oscillations fluctuate because of the variations of the steps shape in $`Q(V_g)`$. On the other hand, if there is strong intersubband mixing in the quantum dot, then the dot charge $`Q`$ undergoes large fluctuations due to coupling to the reservoirs via transmitted subbands. Then quantization of $`Q`$ in the open quantum dot is destroyed, and the charge effects in $`G(V_g)`$ are smeared out. Nevertheless, in submicrometer quantum dots the transverse quantum in constrictions $`\mathrm{}\omega 1`$ meV is usually noticeably greater than the charging energy and pronounced Coulomb oscillations can be observed near and below the conductance treshold where the saw-like dependence of $`V_b(V_g)`$ is retained. ## V Conclusions We have investigated the low-temperature properties of an open quantum dot electrostatically defined by a split gate, and overlaying narrow finger gates at zero magnetic field. Almost periodic and continuous oscillations superimposed upon ballistic conductance steps and Fabry-Pérot resonances are observed even when the conductance through the quantum dot is greater than $`2e^2/h`$. A direct transition of conductance oscillations for $`G>2e^2/h`$ to those for $`G<2e^2/h`$ is observed with decreasing barrier transparencies. The temperature dependence of the observed oscillating features for $`G>2e^2/h`$ and modelling of electron transport excludes the interpretation that they are due to tunneling through single-particle confinement energy states within the dot. Calculated capacitances of the dot to the gates and reservoirs confirm the Coulomb charging nature of the oscillations. Modelling the electrostatics and electron transmission through the quantum dot show that intersubband mixing in our device is greatly reduced in comparison with more standard quantum dots. We have found that in this new design of quantum dot device the charging energy is approximately equal to the subband spacing in the barriers. Suppression of intersubband mixing and high sensitivity of barrier transparency to variations of the Fermi level in the dot made it possible to observe smoothed charged and interference effects over a wide conductance range $`0<G<6e^2/h`$. These results suggest that at zero magnetic field charging effects can occur in the presence of a fully transmitted 1D channel, in contrast to the current experimental and theoretical understanding of Coulomb charging. ###### Acknowledgements. This work was supported by programs of Ministry of Science of Russian Federation “Physics of Solid-State Nanostructures” (grant No. 98-1102) and “Prospective Technologies and Devices for Micro- and Nanoelectronics” (No. 02.04.5.1), and by program “Universities of Russia—Fundamental Research” (No. 1994). The work at Cambridge was funded by the Engineering and Physical Sciences Research Council (EPSRC), United Kingdom. C.T.L. is grateful for financial support from National Sciences Council (grant No. 89-2112-M-002-052). O.A.T., V.A.T. and D.G.B. are grateful to Kvon Ze Don and M. V. Éntin for discussions. O.A.T. and D.G.B. thank the colleagues from Cavendish Lab., especially C. J. B. Ford, for fruitful discussions and hospitality.
warning/0003/gr-qc0003112.html
ar5iv
text
# Homogeneous magnetic fields in fully anisotropic string cosmological backgrounds \[ ## Abstract We present new solutions of the string cosmological effective action in the presence of a homogeneous Maxwell field with pure magnetic component. Exact solutions are derived in the case of space-independent dilaton and vanishing torsion background. In our examples the four dimensional metric is either of Bianchi-type III and VI<sub>-1</sub> or Kantowski-Sachs. \] The solutions of the low-energy string cosmological effective action are, by their nature anisotropic . In this context, it is important to analyze the role of the different possible sources of anisotropy in the low energy string effective action. An interesting and motivated source of anisotropy is represented by pure magnetic fields . Exact string cosmological solutions (containing a magnetic field) were recently found for Bianchi-type I background geometries in the case of a vanishing torsion background and for a space-independent dilaton field . It was shown that while the anisotropy can increase for some time, the string tension corrections (combined with the post big-bang evolution) are likely to force the anisotropy to decay. The purpose of our paper is very simple: we want to generalize the Bianchi-type I solutions of the string cosmological effective action to the case where the homogeneous magnetic field is contained in more general homogeneous (but still anisotropic) metrics. In the context of the low-energy string effective action, the analysis of Bianchi-type classes has been performed (in the absence of any electromagnetic background) by Gasperini and Ricci . In the Bianchi-type I case the homogeneous magnetic (or electric) background is not constrained by the group theoretical properties of the (Abelian) isometry group. In spite of this observation the solutions of the string cosmological effective action are not trivial in the sense that they cannot be found in the usual general relativity literature. The reason is that the dilaton field $`\varphi `$ couples directly to the kinetic term of the Abelian gauge field. In four space-time dimensions such a coupling can be expressed as $`e^\varphi F_{\mu \nu }F^{\mu \nu }`$ where $`\mu ,\nu =0,\mathrm{}3`$ and $`F_{\mu \nu }`$ is the Abelian gauge field strength. Bianchi models (different from Bianchi-type I) are characterized by a non-Abelian isometry group: the Killing vectors (leaving invariant the three-dimensional spatial submanifold) do not commute. Therefore, we face the problem of accommodating an Abelian gauge field in a geometry whose algebraic structure is non-Abelian. The consequence of this remark is that not all non-Abelian isometry groups will be able to support a homogeneous electromagnetic field. More precisely, only Bianchi-type I, II, III, VI<sub>-1</sub> and VII<sub>0</sub> can accommodate a homogeneous magnetic field polarized along a fixed direction . This observation holds in the case of general relativity (where the coupling of the metric to the Maxwell fields is dictated by the equivalence principle) and also in our case provided the dilaton field is space-independent and provided we deal with pure magnetic field (i.e. the homogeneous components of the electric fields are all vanishing). In order to show this point in detail let us look at the relevant Maxwell’s equations. Then we will specify the background geometry which is uniquely characterized by the group theoretical properties of the algebra of the Killing vectors. Finally, we will be able to show which Bianchi-type metrics can (or cannot) be compatible with the presence of a homogeneous (Abelian) magnetic field. Let us start from the Maxwell’s equations and from the related Bianchi identities $$_\mu \left(e^\varphi \sqrt{g}F^{\mu \nu }\right)=0,_\mu \left(\sqrt{g}\stackrel{~}{F}^{\mu \nu }\right)=0,$$ (1) where $`\stackrel{~}{F}^{\mu \nu }`$ is the dual field strength.. In the absence of electric fields and for the case of space-independent dilaton background the gauge field strengths can be written, in the language of differential forms, as $$F=\frac{1}{2}B_iϵ_{ijk}\sigma ^j\sigma ^k,^{}F=B_i\sigma ^i\sigma ^0,$$ (2) where $`\sigma ^0=dt`$ and $`\sigma ^i=R_j^i\omega ^j`$ form the Cartan basis of one-form; the two-form $`F`$ represents the electromagnetic field and $`{}_{}{}^{}F`$ its dual. The matrices $`R_i^j`$ depend only on time. If we take, for instance Cartesian coordinates as local basis the relation between the exterior derivative of the invariant basis is given by : $$d\omega ^i=\frac{1}{2}C_{jk}^i\omega ^j\omega ^k$$ (3) where $`C_{jk}^i`$ are the structure constants of the three dimesional isometry group which appear in the commutation relations of the Killing vectors. For the case of Bianchi-type I we have that $`C_{jk}^i`$ are identically zero (i.e. the isometry group is Abelian). For the other Bianchi-type geometries the structure constants are not all vanishing (i.e. the isometry group is non-Abelian). By now expressing Eqs. (1) in the invariant basis of the $`\omega ^i`$ and by recalling that the dilaton field is only time dependent we obtain a system of three linear equations, namely $`\dot{Z}_a=0,M_jC_{ab}^jϵ_{cab}=0,`$ (4) $`Z_jC_{aj}^j=0,`$ (5) where the over-dot denotes differentiation with respect to $`t`$ and where $`Z_a={\displaystyle \frac{1}{2}}B_iϵ_{ijk}ϵ_{abc}R_b^jR_c^k,`$ (6) $`M_j={\displaystyle \frac{1}{2}}R_j^iB_i.`$ (7) The solution of the first of Eqs. (4) gives $`Z_a=q_a`$ where $`q_a`$ are constant vectors. Therefore Eqs. (4)–(5) can be written as $$Z_a=q_a,q_aC_{ja}^j=0,M_jC_{ab}^j=0.$$ (8) It is clear from these expressions that in the case where all the $`C_{jk}^i=0`$ the magnetic field components are not constrained since all the equations (8) are identically statisfied for any vector $`B_i`$ and $`q_a`$. In the Bianchi-type II case we have that the only non-vanishing structure constant is $`C_{23}^1=1`$. In this case the first two equations of (8) are always satisfied, whereas the third of Eqs. (8) constrains the magnetic field to vanish in one specific direction (i.e. the magnetic field can only have two independent components). Similarly, in the Bianchi-type III case the only non-vansihing structure constant is $`C_{23}^2=1`$ and therefore both, the second and the third equation of (8) will be non trivially satisfied. Thus in Bianchi-type III models the magnetic field will have only one independent component: the other two will have to vanish for the compatibility with the non-Abelian structure of the isometry group. In the Bianchi-type IV and V cases the structure constants are, respectively $`C_{13}^1=C_{23}^1=C_{23}^2=1,\mathrm{IV},`$ (9) $`C_{13}^1=C_{23}^2=1,\mathrm{V}.`$ (10) From the second of Eq. (8) we see that, in the Bianchi-type IV and V, $`q_3=0`$. Moreover form the third of Eq. (8) we see that $`B_k=0`$. Therefore, in the Bianchi-type IV and V the isometry group is only compatible with a vanishing (pure) magnetic field. In the Bianchi-type VI and VII the structure constants are, respectively, $`C_{13}^1=1,C_{23}^2=h,\mathrm{VI},`$ (11) $`C_{32}^1=C_{13}^2=1,C_{23}^2=h,\mathrm{VII}.`$ (12) In the case of Bianchi-type VI a magnetic field can be only present (though with only one independent component) if $`h=1`$ . Only in this case one can consistently satisfy the second of Eqs. (8). In the Bianchi-type VII geometry, for the same reason, a magnetic field (with one independent component) can be present only if $`h=0`$. Finally in the Bianchi-type VIII and IX metrics the structure constants are, respectively $`C_{23}^1=C_{12}^3=C_{13}^2=1,\mathrm{VIII},`$ (13) $`C_{ij}^k=ϵ_{ijk},\mathrm{IX}.`$ (14) In these two last cases we can see that the third equation of (8) implies that $`B_iR^ij=0`$ which is identically satisfied only in the absence of any magnetic background. In conclusion we showed that if the dilaton field only depends upon time and if the electric field is absent only the Bianchi-type I, II, III, VI<sub>-1</sub>, VII<sub>0</sub> are compatible with the presence of a magnetic background. In the Bianchi-type I case no constraint on the components of the magnetic field is present. In the other Bianchi-type models the magnetic field is always constrained. In the Bianchi-type IV, V, VI<sub>h≠1</sub>, VII<sub>h≠0</sub>, VIII, IX cases we cannot accommodate a pure (Abelian) magnetic field. What we discussed, up to now is only a sufficient condition in order to accommodate a magnetic background with time dependent dilaton in a Bianchi-type geometry. Our analysis does not guarantee the existence of explicit (exact) solutions. In the following, as an example, we want to present few exact solutions containing both a magnetic field, a time-dependent dilaton and a background geometry with non-Abelian isometry group. The action <sup>*</sup><sup>*</sup>*We work in string units and we set the string tension $`\alpha ^{}`$ equal to one we ought to study is $$S=d^4x\sqrt{g}e^\varphi \left[R+g^{\alpha \beta }_\alpha \varphi _\beta \varphi +\frac{1}{4}F_{\mu \nu }F^{\mu \nu }\right]$$ (15) whose associated equations of motion can be written as $`Rg^{\alpha \beta }_\alpha \varphi _\beta \varphi +2g^{\alpha \beta }_\alpha _\beta \varphi ={\displaystyle \frac{1}{4}}F_{\alpha \beta }F^{\alpha \beta },`$ (16) $`R_\mu ^\nu +_\mu ^\nu \varphi +{\displaystyle \frac{1}{2}}F_{\mu \alpha }F^{\nu \alpha }=0,`$ (17) $`_\mu \left[e^\varphi F^{\mu \nu }\right]=0.`$ (18) where $`_\mu `$ denotes covariant differentiation with respect to the background metric. Notice that in writing the previous action we assumed zero central charge deficit. This means that the six internal dimensions are assumed to be compactified with constant radius. For sake of simplicity we choose to parametrize our line element as $$ds^2=dt^2a^2(t)dx^2b^2(t)e^{2\lambda x}dy^2c^2(t)dz^2.$$ (19) Notice that for $`\lambda =1`$ we have the Bianchi III line element, whereas the case $`\lambda =2`$ we get the Bianchi VI<sub>-1</sub> case. In the metric of Eq. (19) the components of Eqs. (16)–(18) can be found by direct projection on the spatial vielbein. Care must be taken, however, in solving the generalized Maxwell equation (18). Eq. (18) together with the Bianchi identities imply, in the absence of sources, that a pure magnetic field can be accommodated along the $`z`$ axis so that $`F_{\alpha \beta }F^{\alpha \beta }=2B^2/a^2b^2`$ ($`B`$ is a constant). With these specifications and by defining the shifted time The over-dot denotes derivation with respect to the cosmic time coordinate $`t`$. derivative of the dilaton $`\dot{\overline{\varphi }}=\dot{\varphi }(H+F+G)`$ together with the expansion rates along the three different spatial directions (i.e. $`H=\dot{a}/a`$, $`F=\dot{b}/b`$ and $`G=\dot{c}/c`$) we get to $`2\ddot{\overline{\varphi }}\dot{\overline{\varphi }}^2(H^2+F^2+G^2)+2{\displaystyle \frac{\lambda ^2}{a^2}}+{\displaystyle \frac{B^2}{2a^2b^2}}=0,`$ (20) $`\ddot{\overline{\varphi }}(H^2+F^2+G^2)=0,`$ (21) $`H\dot{\overline{\varphi }}\dot{H}+{\displaystyle \frac{\lambda ^2}{a^2}}+{\displaystyle \frac{B^2}{2a^2b^2}}=0,`$ (22) $`F\dot{\overline{\varphi }}\dot{F}+{\displaystyle \frac{\lambda ^2}{a^2}}+{\displaystyle \frac{B^2}{2a^2b^2}}=0,`$ (23) $`{\displaystyle \frac{\lambda }{a^2}}(HF)=0,`$ (24) In order to solve exactly the previous system we can define a new time coordinate, namely, $`dt=e^{\overline{\varphi }}d\eta `$. By denoting with a prime the derivatives with respect to $`\eta `$ we can re-write the system of equations (20)–(24) as $`𝒢^{}=0,=,`$ (25) $`^{}=\lambda ^2a^2c^2e^{2\varphi }+{\displaystyle \frac{B^2}{2}}c^2e^{2\varphi },`$ (26) $`\varphi ^{\prime \prime }+\varphi _{}^{}{}_{}{}^{2}2Q\varphi ^{}Q^{}+P=0,`$ (27) $`Q(\eta )={\displaystyle \frac{d\mathrm{ln}\sqrt{g}}{d\eta }},P=2(2+𝒢)`$ (28) where the prime denotes now the derivative with respect to $`\eta `$ and where $`=a^{}/a`$, $`=b^{}/b`$, $`𝒢=c^{}/c`$. By linearly combining the previous equations and by defining $`\mathrm{\Phi }=\varphi \mathrm{ln}c`$ the equation for $`\mathrm{\Phi }`$ $$\mathrm{\Phi }^{\prime \prime }=\frac{B^2}{2}e^{2\mathrm{\Phi }},$$ (29) can be integrated directly twice with the result that $$\mathrm{\Phi }=\mathrm{ln}\left[\frac{\delta }{\gamma }\mathrm{cosh}\gamma (\eta \eta _0)\right],$$ (30) where $`\delta =B/\sqrt{2}`$; $`\gamma `$ and $`\eta _0`$ are integration constants. By now going to Eq. (26) and by making the ansatz $`a(\eta )=e^\mathrm{\Phi }f(\eta )`$ we get an equation for $`f(\eta )`$ $$\left(\frac{f^{}}{f}\right)^{}=\lambda ^2f^2,$$ (31) which, again, can be directly integrated with the result that $$f(\eta )=\frac{\beta }{\lambda |\mathrm{sinh}\beta (\eta \eta _0)|}$$ (32) Putting all our results together we obtain that $`\varphi (\eta )=\mathrm{ln}c_0+𝒢_0\eta +\mathrm{ln}\left[{\displaystyle \frac{\delta }{\gamma }}\mathrm{cosh}\gamma (\eta \eta _0)\right],`$ (33) $`a(\eta )={\displaystyle \frac{\beta \delta }{\lambda \gamma }}{\displaystyle \frac{\mathrm{cosh}\gamma (\eta \eta _0)}{|\mathrm{sinh}\beta (\eta \eta _0)|}},`$ (34) $`c(\eta )=c_0e^{𝒢_0\eta }.`$ (35) By requiring the consistency of this solution with the constraint equation we find that $`2\beta ^2=\gamma ^2+𝒢_0^2`$. It is interesting to notice that the solution we just obtained can be related to the solution of the same set of Eqs. (16)–(18) in a Kantowski-Sachs metric with a magnetic field oriented along the radial coordinate. The Kantowski-Sachs line element can be written as $$ds^2=dt^2m^2(t)dr^2n^2(t)[d\theta ^2+A^2(\theta )d\chi ^2].$$ (36) where $`A(\theta )=\mathrm{sin}\theta ,\theta ,\mathrm{sinh}\theta `$. If a magnetic field directed along the radial direction is present, then, the equations of motion (17)–(18) will become $`^{}=0,`$ (37) $`𝒩^{}=kn^2m^2e^{2\varphi }+{\displaystyle \frac{B^2}{2}}m^2e^{2\varphi },`$ (38) $`\varphi ^{\prime \prime }+\varphi _{}^{}{}_{}{}^{2}2Q\varphi ^{}Q^{}+P=0,`$ (39) $`Q(\eta )=(+2𝒩),P=2(𝒩+2𝒩^2),`$ (40) where $`k=+1,0,1`$ if, respectively, $`A(\theta )=\mathrm{sin}\theta ,\theta ,\mathrm{sinh}\theta `$. This system can be integrated with the same techniques we discussed in the Bianchi case. Notice, moreover, that in the case $`k=1`$ the analogy with the solutions (33)–(35) is complete. We want now to discuss the cosmic-time evolution of our solutions. Let us focus, for instance, on the Bianchi-type II case ($`\lambda =1`$). On the basis of our results we can consistently choose $`\beta =\gamma =1`$. The consistency relation $`2\beta ^2=\gamma ^2+𝒢_0^2`$ implies that $`𝒢_0=1`$ in string units. The relation between $`\eta `$ and the cosmic time coordinate can be obtained by direct integration of $`a^2ce^\varphi d\eta =dt`$. The result is $$\eta (t)=\mathrm{ln}\left[\frac{\delta }{t}+\sqrt{1+\frac{\delta ^2}{t^2}}\right].$$ (41) We want, in particular, to analyze our solutions for $`t<0`$. The solutions given in Eqs. (33)–(35) can then be expressed as $`a(t)={\displaystyle \frac{\sqrt{t^2+\delta ^2}}{\delta }},`$ (42) $`c(t)=\left({\displaystyle \frac{\delta }{t}}+\sqrt{{\displaystyle \frac{\delta ^2}{t^2}}+1}\right)^1,`$ (43) $`\varphi (t)=\mathrm{ln}\left[{\displaystyle \frac{\sqrt{1+\frac{\delta ^2}{t^2}}}{\frac{\delta }{t}+\sqrt{1+\frac{\delta ^2}{t^2}}}}\right].`$ (44) In order to write the solutions in this form we also have chosen $`c_0=1`$ and $`\eta _0=0`$. The choice $`\eta _0=0`$ simply translates in the origin the curvature singularity which appears clearly in $`F=\dot{c}/c`$. For $`t<0`$ the dilaton coupling grows. The scale factor $`a(t)`$ contracts to a minimal value determined by $`\delta `$, i.e. by the value of the magnetic field in string units. The scale factor $`c(t)`$ expands towards a singularity at $`t=0`$. It is finally interesting to compute the shear parameter in the case we just described. The shear parameter measures the degree of isotropization of a given anisotropic solution . We define the shear parameter, in our case , as $$r(t)=\frac{HG}{H+2G}.$$ (45) Using eqs. (42)–(43) we find that $$r(t)=\frac{\sqrt{t^6+\delta ^2t^4}\delta ^3\delta t^2}{2\sqrt{t^6+\delta ^2t^4}+\delta ^3+\delta t^2}.$$ (46) By expanding the above expression in Taylor series around $`t=0_{}`$ we have that $$r(t)=1+\frac{\delta +2}{\delta ^2}t^2+𝒪(t^3).$$ (47) namely we see that the share parameter grows, a behavior similar to the one discussed in the Bianchi-type I case. This behavior is then not surprising and it is due to the fact that, in our discussion, we did not take into account the string tension corrections to the low-energy action. If string tension corrections are (naively) included the shear parameter is forced to decrease. In conclusion we discussed the possible inclusion of a pure (Abelian) magnetic field in the low-energy string effective action. We found that, in the absence of any torsion background, a magnetic field and a time-dependent dilaton field can be simultaneously accommodated in Bianchi-type I, II, III, VI<sub>-1</sub>, VII<sub>0</sub> backgrounds. As an example of our considerations we found new exact solutions of the low-energy string effective action in the Bianchi-type III, $`VI_1`$ and Kantowski-Sachs cases. We also discussed the physical properties of the obtained solutions and we argued that they are not qualitatively different from the ones deduced in the Bianchi-type I background .
warning/0003/hep-th0003273.html
ar5iv
text
# References 1. Introduction. During last few years there was a considerable interest in applying the general method of nonlinear realizations to systems with partial breaking of global supersymmetries (PBGS), first of all to the superbranes as a notable example of such systems (see, e.g., and refs. therein). On this path one meets two problems. The first one is purely computational. Following the general prescriptions of nonlinear realizations, one is led to include into the coset, alongside with the spontaneously broken translation and supertranslation generators, also the appropriate part of generators of the automorphism group for the given supersymmetry algebra (including those of the Lorentz group). This makes the computations beyond the linearized approximation rather complicated. Moreover, sometimes these additional symmetries which we should take into account at the step of doing the coset routine appear to be explicitly broken at the level of the invariant action (see, e.g., refs. ), with no clear reasons for this. The second, closely related difficulty is lacking of a systematic procedure for constructing the PBGS actions. In all the cases elaborated so far, the PBGS Lagrangians cannot be constructed in a manifestly invariant way from the relevant Cartan forms: under the broken supersymmetry transformations they are shifted by the spinor or $`x`$-derivatives (like the WZNW or Chern-Simons Lagrangians). In the present note we argue, on several instructive examples, that the automorphism symmetries can be ignored if we are interested only in the equations of motion for the given PBGS system. This radically simplifies the calculations, resulting in rather simple manifestly covariant equations in which all nonlinearities are hidden inside the covariant derivatives. 2. $`N=1,D=4`$ supermembrane and D2-brane. To clarify the main idea of our approach, let us start from the well known systems with partially broken global supersymmetries . Our goal is to get the corresponding superfield equations of motion in terms of the worldvolume superfields starting from the nonlinear realization of the global supersymmetry group. The supermembrane in $`D=4`$ spontaneously breaks half of four $`N=1,D=4`$ supersymmetries and one translation. Let us split the set of generators of $`N=1D=4`$ Poincaré superalgebra (in the $`d=3`$ notation) into the unbroken $`\{Q_a,P_{ab}\}`$ and broken $`\{S_a,Z\}`$ ones ($`a,b=1,2`$). The $`d=3`$ translation generator $`P_{ab}=P_{ba}`$ together with the generator $`Z`$ form the $`D=4`$ translation generator. The basic anticommutation relations read <sup>1</sup><sup>1</sup>1Hereafter, we consider the spontaneously broken supersymmetry algebras modulo possible extra central-charge type terms which should be present in the full algebra of the corresponding Noether currents to evade the no-go theorem of ref. along the lines of ref. . $$\{Q_a,Q_b\}=P_{ab},\{Q_a,S_b\}=ϵ_{ab}Z,\{S_a,S_b\}=P_{ab}.$$ (1) In contrast to our previous considerations , here we prefer to construct the nonlinear realization of the superalgebra (1) itself, ignoring all generators of the automorphisms of (1) (the spontaneously broken as well as unbroken ones), including those of $`D=4`$ Lorentz group $`SO(1,3)`$. Thus, we put all generators into the coset and associate the $`N=1,d=3`$ superspace coordinates $`\{\theta ^a,x^{ab}\}`$ with $`Q_a,P_{ab}`$. The remaining coset parameters are Goldstone superfields, $`\psi ^a\psi ^a(x,\theta ),qq(x,\theta )`$. A coset element $`g`$ is defined by $$g=e^{x^{ab}P_{ab}}e^{\theta ^aQ_a}e^{qZ}e^{\psi ^aS_a}.$$ (2) As the next step of the coset formalism, one constructs the Cartan 1-forms $$g^1dg=\omega _Q^aQ_a+\omega _P^{ab}P_{ab}+\omega _ZZ+\omega _S^aS_a,$$ (3) $`\omega _Z`$ $`=`$ $`dq+\psi _ad\theta ^a,\omega _P^{ab}=dx^{ab}+{\displaystyle \frac{1}{4}}\theta ^{(a}d\theta ^{b)}+{\displaystyle \frac{1}{4}}\psi ^{(a}d\psi ^{b)},`$ $`\omega _Q^a`$ $`=`$ $`d\theta ^a,\omega _S^a=d\psi ^a;.`$ (4) and define the covariant derivatives $$𝒟_{ab}=(E^1)_{ab}^{cd}_{cd},𝒟_a=D_a+\frac{1}{2}\psi ^bD_a\psi ^c𝒟_{bc}=D_a+\frac{1}{2}\psi ^b𝒟_a\psi ^c_{bc},$$ (5) where $`D_a={\displaystyle \frac{}{\theta ^a}}+{\displaystyle \frac{1}{2}}\theta ^b_{ab},\{D_a,D_b\}=_{ab},`$ (6) $`E_{ab}^{cd}={\displaystyle \frac{1}{2}}(\delta _a^c\delta _b^d+\delta _a^d\delta _b^c)+{\displaystyle \frac{1}{4}}(\psi ^c_{ab}\psi ^d+\psi ^d_{ab}\psi ^c).`$ (7) They obey the following algebra $`[𝒟_{ab},𝒟_{cd}]=𝒟_{ab}\psi ^f𝒟_{cd}\psi ^g𝒟_{fg},`$ $`[𝒟_{ab},𝒟_c]=𝒟_{ab}\psi ^f𝒟_c\psi ^g𝒟_{fg},`$ $`\{𝒟_a,𝒟_b\}=𝒟_{ab}+𝒟_a\psi ^f𝒟_b\psi ^g𝒟_{fg}.`$ (8) Not all of the above Goldstone superfields $`\{q(x,\theta ),\psi ^a(x,\theta )\}`$ must be treated as independent. Indeed, $`\psi _a`$ appears inside the form $`\omega _Z`$ linearly and so can be covariantly eliminated by the manifestly covariant constraint (inverse Higgs effect ) $$\omega _Z|_{d\theta }=0\psi _a=𝒟_aq,$$ (9) where $`|_{d\theta }`$ means the ordinary $`d\theta `$-projection of the form. Thus the superfield $`q(x,\theta )`$ is the only essential Goldstone superfield needed to present the partial spontaneous breaking $`N=1,D=4N=1,d=3`$ within the coset scheme. Now we are ready to put additional, manifestly covariant constraints on the superfield $`q(x,\theta )`$, in order to get dynamical equations. The main idea is to covariantize the “flat” equations of motion. Namely, we simply replace the flat covariant derivatives in the standard equation of motion for the bosonic scalar superfield in $`d=3`$ $$D^aD_aq=0$$ (10) by the covariant ones (5) $$𝒟^a𝒟_aq=0.$$ (11) The equation (11) coincides with the equation of motion of the supermembrane in $`D=4`$ as it was presented in . Thus, we conclude that, at least in this specific case, additional superfields-parameters of the extended coset with all the automorphism symmetry generators included are auxiliary and can be dropped out if we are interested in the equations of motion only. Actually, in eq. (11) was deduced, proceeding from the $`D=4`$ Lorentz covariant coset formalism with preserving all initial symmetries. This means that (11), having been now reproduced from the coset involving only the translations and supertranslations generators, possesses the hidden covariance under the full $`D=4`$ Lorentz group. On the other hand, one more automorphism symmetry of the $`N=1,D=4`$ supersymmetry algebra, “$`\gamma _5`$” symmetry, is explicitly broken in eq. (11), and there is no way to keep it. In the $`d=3`$ notation this symmetry is realized as an extra $`SO(2)`$ with respect to which the generators $`Q_a`$ and $`S_b`$ and, respectively, the coset parameters $`\theta ^a,\psi ^a`$ form a 2-vector. This symmetry is spontaneously broken at the level of the transformation laws, with the auxiliary field of $`q(x,\theta )`$ being the relevant Goldstone field. From eq. (11) we conclude that it cannot be preserved even in this spontaneously broken form when $`q`$ is subjected to the dynamical equation: one can preserve the spontaneously broken $`D=4`$ Lorentz symmetry at most. This $`U(1)`$ is explicitly broken in the off-shell PBGS action of ref. , as well as in the corresponding Green-Schwarz action . A similar phenomenon was observed in refs. for the $`N=(1,0),D=6`$ 3-brane. There, the auxiliary fields of the basic worldvolume $`N=1,d=4`$ Goldstone chiral supermultiplet are the Goldstone fields parameterizing the coset $`SU(2)_A/U(1)_A`$ of the automorphism $`SU(2)_A`$ group of $`N=(1,0),D=6`$ Poincaré superalgebra, and the coset part of $`SU(2)_A`$ is realized as nonlinear shifts of these fields. In the superfield equations of motion of the 3-brane and the corresponding off-shell action this $`SU(2)_A`$ is explicitly broken down to $`U(1)_A`$, though the spontaneously broken $`D=6`$ Lorentz symmetry is still preserved. As a straightforward application of the idea that the automorphism symmetries are irrelevant when deducing the equations of motion, let us consider the case of the “space-time filling” D2-brane (i.e. having no scalar fields in its worldvolume multiplet the field content of which is that of $`N=1,d=3`$ vector multiplet). The main problem with the description of D-branes within the standard nonlinear realization approach is the lack of the coset generators to which one could relate the gauge fields as the coset parameters <sup>2</sup><sup>2</sup>2For the covariant field strengths as Goldstone fields such generators can still be found in the automorphism symmetry algebras .. So we do not know how interpret the gauge fields as coset parameters in this case <sup>3</sup><sup>3</sup>3It seems that the existing interpretation of gauge fields as the coset fields can be generalized to the PBGS case only on the way of non-trivial unification of the gauge group algebra with that of supersymmetry, so that the gauge group transformations appeared in the closure of supersymmetries before any gauge-fixing as a sort of tensorial central charges.. Let us show how these difficulties can be circumvented in the present approach. The superalgebra we start with is the same algebra (1), but now without the central charge $$Z=0.$$ The coset element $`g`$ contains only one Goldstone superfield $`\psi ^a`$ which now must be treated as the essential one, and the covariant derivatives coincide with (5). Bearing in mind to end up with the irreducible field content of $`N=1,d=3`$ vector multiplet, we are led to treat $`\psi ^a`$ as the corresponding superfield strength and to find the appropriate covariantization of the flat irreducibility constraint and the equation of motion. In the flat case the $`d=3`$ vector multiplet is represented by a $`N=1`$ spinor superfield strength $`\mu _a`$ subjected to the Bianchi identity : $$D^a\mu _a=0\left\{\begin{array}{c}D^2\mu _a=_{ab}\mu ^b,\hfill \\ _{ab}D^a\mu ^b=0.\hfill \end{array}\right\}.$$ This leaves in $`\mu _a`$ the first fermionic (Goldstone) component, together with the divergenceless vector $`F_{ab}D_a\mu _b|_{\theta =0}`$ (i.e., just the gauge field strength). The equation of motion reads $$D^2\mu _a=0.$$ (12) In accordance with our approach, we propose the following equations which should describe the D2-brane: $$(a)𝒟^a\psi _a=0,(b)𝒟^2\psi _a=0.$$ (13) The equation $`(a)`$ is a covariantization of the irreducibility constraint (S0.Ex6) while $`(b)`$ is the covariant equation of motion. In order to see which kind of dynamics is encoded in (13), we considered it in the bosonic limit. We found that it amounts to the following equations for the vector $`V_{ab}𝒟_a\psi _b|_{\theta =0}`$: $$\left(_{ac}+V_a^mV_c^n_{mn}\right)V_b^c=0.$$ (14) One can wonder how these nonlinear but polynomial equations can be related to the nonpolynomial Born-Infeld theory which is just the bosonic core of the superfield D2-brane theory as was explicitly demonstrated in . The trick is to rewrite the parts of the equation (14), respectively antisymmetric and symmetric in the indices $`\{a,b\}`$, as follows: $`_{ab}\left({\displaystyle \frac{V^{ab}}{2V^2}}\right)=0,`$ (15) $`_{ac}\left({\displaystyle \frac{V_b^c}{2+V^2}}\right)+_{bc}\left({\displaystyle \frac{V_a^c}{2+V^2}}\right)=0,`$ (16) where $`V^2V^{mn}V_{mn}`$. After passing to the “genuine” field strength $$F^{ab}=\frac{2V^{ab}}{2V^2}_{ab}F^{ab}=0,$$ (17) the equation of motion (16) takes the familiar Born-Infeld form $$_{ac}\left(\frac{F_b^c}{\sqrt{1+2F^2}}\right)+_{bc}\left(\frac{F_a^c}{\sqrt{1+2F^2}}\right)=0.$$ (18) Thus we have proved that the bosonic part of our system (13) indeed coincides with the Born-Infeld equations. One may explicitly show that the full equations (13) are equivalent to the worldvolume superfield equation following from the off-shell D2-brane action given in (augmented with the Bianchi identity (S0.Ex6)). An indirect proof is based on the fact that (13) is an $`N=1`$ extension of the bosonic $`d=3`$ Born-Infeld equations, such that it possesses one more nonlinearly realized supersymmetry completing the explicit one to $`N=2,d=3`$ superalgebra (1) with $`Z=0`$. On the other hand, the $`N=1,d=3`$ superfield action of is uniquely specified by requiring it to possess this second supersymmetry. Hence both types of equations should be equivalent. In closing this Section, it is worth mentioning that the equations (14) which equivalently describe the bosonic Born-Infeld dynamics in $`d=3`$, look much simpler than the standard ones (17), (18). 3. D3-brane. As another interesting application of the proposed approach, we shall consider the space-time filling D3-brane in $`d=4`$. This system amounts to the PBGS pattern $`N=2,d=4N=1,d=4`$, with a nonlinear generalization of $`N=1,d=4`$ vector multiplet as the Goldstone multiplet . The off-shell superfield action for this system and the related equations of motion are known , but the latter have never been derived directly from the coset approach. Our starting point is the $`N=2,d=4`$ Poincaré superalgebra without central charges: $$\{Q_\alpha ,\overline{Q}_{\dot{\alpha }}\}=2P_{\alpha \dot{\alpha }},\{S_\alpha ,\overline{S}_{\dot{\alpha }}\}=2P_{\alpha \dot{\alpha }}.$$ (19) Assuming the $`S_\alpha ,\overline{S}_{\dot{\alpha }}`$ supersymmetries to be spontaneously broken, we introduce the Goldstone superfields $`\psi ^\alpha (x,\theta ,\overline{\theta }),\overline{\psi }^{\dot{\alpha }}(x,\theta ,\overline{\theta })`$ as the corresponding parameters in the following coset (we use the same notation as in ) $$g=e^{ix^{\alpha \dot{\alpha }}P_{\alpha \dot{\alpha }}}e^{i\theta ^\alpha Q_\alpha +i\overline{\theta }_{\dot{\alpha }}\overline{Q}^{\dot{\alpha }}}e^{i\psi ^\alpha S_\alpha +i\overline{\psi }_{\dot{\alpha }}\overline{S}^{\dot{\alpha }}}.$$ (20) With the help of the Cartan forms $`g^1dg`$ $`=`$ $`i\omega ^{\alpha \dot{\alpha }}P_{\alpha \dot{\alpha }}+i\omega _Q^\alpha Q_\alpha +i\overline{\omega }_{Q\dot{\alpha }}\overline{Q}^{\dot{\alpha }}+i\omega _S^\alpha S_\alpha +i\overline{\omega }_{S\dot{\alpha }}\overline{S}^{\dot{\alpha }},`$ $`\omega ^{\alpha \dot{\alpha }}`$ $`=`$ $`dx^{\alpha \dot{\alpha }}i\left(\theta ^\alpha d\overline{\theta }^{\dot{\alpha }}+\overline{\theta }^{\dot{\alpha }}d\theta ^\alpha +\psi ^\alpha d\overline{\psi }^{\dot{\alpha }}+\overline{\psi }^{\dot{\alpha }}d\psi ^\alpha \right),`$ $`\omega _Q^\alpha `$ $`=`$ $`d\theta ^\alpha ,\overline{\omega }_Q^{\dot{\alpha }}=d\overline{\theta }^{\dot{\alpha }},\omega _S^\alpha =d\psi ^\alpha ,\overline{\omega }_S^{\dot{\alpha }}=d\overline{\psi }^{\dot{\alpha }},`$ (21) one can define the covariant derivatives $`𝒟_{\alpha \dot{\alpha }}`$ $`=`$ $`\left(E^1\right)_{\alpha \dot{\alpha }}^{\beta \dot{\beta }}_{\beta \dot{\beta }},`$ $`𝒟_\alpha `$ $`=`$ $`D_\alpha i\left(\overline{\psi }^{\dot{\beta }}D_\alpha \psi ^\beta +\psi ^\beta D_\alpha \overline{\psi }^{\dot{\beta }}\right)𝒟_{\beta \dot{\beta }},`$ $`\overline{𝒟}_{\dot{\alpha }}`$ $`=`$ $`\overline{D}_{\dot{\alpha }}i\left(\overline{\psi }^{\dot{\beta }}\overline{D}_{\dot{\alpha }}\psi ^\beta +\psi ^\beta \overline{D}_{\dot{\alpha }}\overline{\psi }^{\dot{\beta }}\right)𝒟_{\beta \dot{\beta }},`$ (22) where $$E_{\alpha \dot{\alpha }}^{\beta \dot{\beta }}=\delta _\alpha ^\beta \delta _{\dot{\alpha }}^{\dot{\beta }}i\psi ^\beta _{\alpha \dot{\alpha }}\overline{\psi }^{\dot{\beta }}i\overline{\psi }^{\dot{\beta }}_{\alpha \dot{\alpha }}\psi ^\beta ,$$ (23) and the flat covariant derivatives are defined as follows $$D_\alpha =\frac{}{\theta ^\alpha }i\overline{\theta }^{\dot{\alpha }}_{\alpha \dot{\alpha }},\overline{D}_{\dot{\alpha }}=\frac{}{\overline{\theta }^{\dot{\alpha }}}+i\theta ^\alpha _{\alpha \dot{\alpha }}.$$ (24) Now we are ready to write the covariant version of the constraints on $`\psi ^\alpha ,\overline{\psi }^{\dot{\alpha }}`$ which define the superbrane generalization of $`N=1,d=4`$ vector multiplet, together with the covariant equations of motion for this system. As is well-known , the $`N=1,d=4`$ vector multiplet is described by a chiral $`N=1`$ field strength $`W_\alpha `$, $$\overline{D}_{\dot{\alpha }}W_\alpha =0,D_\alpha \overline{W}_{\dot{\alpha }}=0,$$ (25) which satisfies the irreducibility constraint (Bianchi identity) $$D^\alpha W_\alpha +\overline{D}_{\dot{\alpha }}\overline{W}^{\dot{\alpha }}=0.$$ (26) The free equations of motion for the vector multiplet read $$D^\alpha W_\alpha \overline{D}_{\dot{\alpha }}\overline{W}^{\dot{\alpha }}=0.$$ (27) It was shown in that the chirality constraints (25) can be directly covariantized $$\overline{𝒟}_{\dot{\alpha }}\psi _\alpha =0,𝒟_\alpha \overline{\psi }_{\dot{\alpha }}=0.$$ (28) These conditions are compatible with the algebra of the covariant derivatives (S0.Ex9). This algebra, with the constraints (28) taken into account, reads $`\{𝒟_\alpha ,𝒟_\beta \}=\{\overline{𝒟}_{\dot{\alpha }},\overline{𝒟}_{\dot{\beta }}\}=0,`$ $`\{𝒟_\alpha ,\overline{𝒟}_{\dot{\beta }}\}=2i𝒟_{\alpha \dot{\beta }}2i(𝒟_\alpha \psi ^\gamma \overline{𝒟}_{\dot{\beta }}\overline{\psi }^{\dot{\gamma }})𝒟_{\gamma \dot{\gamma }},`$ $`\{𝒟_\alpha ,𝒟_{\gamma \dot{\gamma }}\}=2i(𝒟_\alpha \psi ^\beta 𝒟_{\gamma \dot{\gamma }}\overline{\psi }^{\dot{\beta }})𝒟_{\beta \dot{\beta }}.`$ (29) The first two relations in (29) guarantee the consistency of the above nonlinear version of $`N=1,d=4`$ chirality. They also imply, like in the flat case, $$(𝒟)^3=(\overline{𝒟})^3=0.$$ (30) The second flat irreducibility constraint, eq. (26), is not so simple to covariantize. The straightforward generalization of (26), $$𝒟^\alpha \psi _\alpha +\overline{𝒟}_{\dot{\alpha }}\overline{\psi }^{\dot{\alpha }}=0,$$ (31) is contradictory. Let us apply the square $`(𝒟)^2`$ to the left-hand side of (31). When hitting the first term in the sum, it yields zero in virtue of the property (30). However, it is not zero on the second term. To compensate for the resulting non-vanishing terms, and thus to achieve compatibility with the algebra (29) and its corollaries (30), one should modify (31) by some higher-order coorrections . Let us argue that the constraints (26) together with the equations of motion (27) can be straightforwardly covariantized as $$𝒟^\alpha \psi _\alpha =0,\overline{𝒟}_{\dot{\alpha }}\overline{\psi }^{\dot{\alpha }}=0.$$ (32) Firstly, we note that no difficulties of the above kind related to the compatibility with the algebra (29) arise on the shell of eqs. (32). As a consequence of (32) and the first two relations in (29) we get $$𝒟^2\psi _\alpha =0,\overline{𝒟}^2\overline{\psi }_{\dot{\alpha }}=0.$$ (33) This set is a nonlinear version of the well-known reality condition and the equation of motion for the auxiliary field of vector multiplet. Then, applying, e.g., $`𝒟_\alpha `$ to the second equation in (32) and making use of the chirality condition (28), we obtain the nonlinear version of the equation of motion for photino $$𝒟_{\alpha \dot{\alpha }}\overline{\psi }^{\dot{\alpha }}(𝒟_\alpha \psi ^\gamma \overline{𝒟}_{\dot{\alpha }}\overline{\psi }^{\dot{\gamma }})𝒟_{\gamma \dot{\gamma }}\overline{\psi }^{\dot{\alpha }}=0.$$ (34) Acting on this equation by one more $`𝒟_\alpha `$ and taking advantage of the equations (33), we obtain: $$[𝒟^\alpha ,𝒟_{\alpha \dot{\alpha }}]\overline{\psi }^{\dot{\alpha }}𝒟_\alpha \psi ^\gamma \{𝒟^\alpha ,\overline{𝒟}_{\dot{\alpha }}\}\overline{\psi }^{\dot{\gamma }}𝒟_{\gamma \dot{\gamma }}\overline{\psi }^{\dot{\alpha }}𝒟_\alpha \psi ^\gamma \overline{𝒟}_{\dot{\alpha }}\overline{\psi }^{\dot{\gamma }}[𝒟^\alpha ,𝒟_{\gamma \dot{\gamma }}]\overline{\psi }^{\dot{\alpha }}=0.$$ (35) After substituting the explicit expressions for the (anti)commutators from (29), we observe that (35) is satisfied identically, i.e. it does not imply any further restrictions on $`\psi ^\alpha ,\overline{\psi }^{\dot{\alpha }}`$. It can be also explicitly checked, in a few lowest orders in $`\psi ^\alpha ,\overline{\psi }^{\dot{\alpha }}`$, that the higher-order corrections to (31) found in are vanishing on the shell of eqs. (32). Thus the full set of equations describing the dynamics of the D3-brane supposedly consists of the generalized chirality constraint (28) and the equations (32). To prove its equivalence to the $`N=1`$ superfield description of D3-brane proposed in , recall that the latter is the $`N=1`$ supersymmetrization of the $`d=4`$ Born-Infeld action with one extra nonlinearly realized $`N=1`$ supersymmetry. So, let us consider the bosonic part of the proposed set of equations. Our superfields $`\psi ,\overline{\psi }`$ contain the following bosonic components: $$V^{\alpha \beta }=V^{\beta \alpha }𝒟^\alpha \psi ^\beta |_{\theta =0},\overline{V}^{\dot{\alpha }\dot{\beta }}=\overline{V}^{\dot{\beta }\dot{\alpha }}\overline{𝒟}^{\dot{\alpha }}\overline{\psi }^{\dot{\beta }}|_{\theta =0},$$ (36) which, owing to (32), obey the following simple equations $$_{\alpha \dot{\alpha }}V^{\alpha \beta }V_\alpha ^\gamma \overline{V}_{\dot{\alpha }}^{\dot{\gamma }}_{\gamma \dot{\gamma }}V^{\alpha \beta }=0,_{\alpha \dot{\alpha }}\overline{V}^{\dot{\alpha }\dot{\beta }}V_\alpha ^\gamma \overline{V}_{\dot{\alpha }}^{\dot{\gamma }}_{\gamma \dot{\gamma }}\overline{V}^{\dot{\alpha }\dot{\beta }}=0.$$ (37) Like in the D2-brane case, in the equations (37) nothing reminds us of the Born-Infeld equations. Nevertheless, it is possible to rewrite these equations in the standard Born-Infeld form. The first step is to rewrite eqs.(37) as $`(1{\displaystyle \frac{1}{4}}V^2\overline{V}{}_{}{}^{2})_{\beta \dot{\alpha }}V_\alpha ^\beta +{\displaystyle \frac{1}{4}}\overline{V}{}_{}{}^{2}V_{\alpha }^{\beta }_{\beta \dot{\alpha }}V^2+{\displaystyle \frac{1}{2}}\overline{V}{}_{\dot{\alpha }}{}^{\dot{\beta }}_{\alpha \dot{\beta }}^{}V^2=0,`$ (38) $`(1{\displaystyle \frac{1}{4}}V^2\overline{V}{}_{}{}^{2})_{\alpha \dot{\beta }}\overline{V}{}_{\dot{\alpha }}{}^{\dot{\beta }}+{\displaystyle \frac{1}{4}}V^2\overline{V}{}_{\dot{\alpha }}{}^{\dot{\beta }}_{\alpha \dot{\beta }}^{}\overline{V}{}_{}{}^{2}+{\displaystyle \frac{1}{2}}V_\alpha ^\beta _{\dot{\alpha }\beta }\overline{V}{}_{}{}^{2}=0.`$ (39) After some algebra, one can bring them into the following equivalent form $$_{\beta \dot{\alpha }}\left(fV_\alpha ^\beta \right)_{\alpha \dot{\beta }}\left(\overline{f}\overline{V}_{\dot{\alpha }}^{\dot{\beta }}\right)=0,_{\beta \dot{\alpha }}\left(gV_\alpha ^\beta \right)+_{\alpha \dot{\beta }}\left(\overline{g}\overline{V}_{\dot{\alpha }}^{\dot{\beta }}\right)=0,$$ (40) where $$f=\frac{\overline{V}{}_{}{}^{2}2}{1\frac{1}{4}V^2\overline{V}^2},g=\frac{\overline{V}{}_{}{}^{2}+2}{1\frac{1}{4}V^2\overline{V}^2}.$$ (41) After introducing the “genuine” field strengths $$F_\alpha ^\beta \frac{1}{2\sqrt{2}}fV_\alpha ^\beta ,\overline{F}_{\dot{\alpha }}^{\dot{\beta }}\frac{1}{2\sqrt{2}}\overline{f}\overline{V}_{\dot{\alpha }}^{\dot{\beta }},$$ (42) first of eqs. (40) is recognized as the Bianchi identity $$_{\beta \dot{\alpha }}F_\alpha ^\beta _{\alpha \dot{\beta }}\overline{F}_{\dot{\alpha }}^{\dot{\beta }}=0,$$ (43) while the second one acquires the familiar form of the Born-Infeld equation $`_{\beta \dot{\alpha }}\left({\displaystyle \frac{1+F^2\overline{F}^2}{\sqrt{(F^2\overline{F}{}_{}{}^{2})^22(F^2+\overline{F}{}_{}{}^{2})+1}}}F_\alpha ^\beta \right)`$ $`+_{\alpha \dot{\beta }}\left({\displaystyle \frac{1F^2+\overline{F}^2}{\sqrt{(F^2\overline{F}{}_{}{}^{2})^22(F^2+\overline{F}{}_{}{}^{2})+1}}}\overline{F}_{\dot{\alpha }}^{\dot{\beta }}\right)=0.`$ (44) Thus, in this new basis the action for our bosonic system is the Born-Infeld action: $$S=d^4x\sqrt{(F^2\overline{F}{}_{}{}^{2})^22(F^2+\overline{F}{}_{}{}^{2})+1}.$$ (45) Now the equivalence of the system (32) to the equations corresponding to the action of ref. , like in the D2-brane case, can be established proceeding from the following two arguments: (i) It is $`N=1`$ supersymmetrization of the $`d=4`$ Born-Infeld equations; (ii) It possesses the second hidden nonlinearly realized supersymmetry lifting $`N=1,d=4`$ to $`N=2,d=4`$. The action given in provides the unique extension of the $`d=4`$ Born-Infeld action with both these requirements satisfied. Hence, both representations should be equivalent to each other. Note that at the full superfield level the redefinition (42) should correspond to passing from the Goldstone fermions $`\psi _\alpha `$, $`\overline{\psi }_{\dot{\alpha }}`$ which have the simple transformation properties in the nonlinear realization of $`N=1,d=4`$ supersymmetry but obey the nonlinear irreducibility constraints, to the ordinary Maxwell superfield strength $`W_\alpha ,\overline{W}_{\dot{\alpha }}`$ defined by eqs. (25), (26). The nonlinear action in was written just in terms of this latter object. The equivalent form (32) of the equations of motion and Bianchi identity is advantageous in that it is manifestly covariant under the second (hidden) supersymmetry, being constructed out of the covariant objects. 4. Conclusions. In this Letter we demonstrated that in many cases one can simplify the analysis of the equations of motion which follow from the coset approach by taking no account of the automorphism group at all. We showed that the equations of motion for the $`N=1,D=4`$ supermembrane, D2- and D3-branes in a flat background have a very simple form when written in terms of Goldstone superfields of nonlinear realizations and the corresponding nonlinear covariant derivatives. As a by-product, we got a new simple form for the $`d=3`$ and $`d=4`$ Born-Infeld theory equations of motion combined with the appropriate Bianchi identities. The remarkable property of this representation is that it involves only a third order nonlinearity in the gauge field strength. Note that the idea to use the geometric and symmetry principles to derive the dynamical equations is not new, of course. For instance, the completely integrable $`d=2`$ equations admit the geometrical interpretation as the vanishing of some curvatures. In the superembedding approach (see and refs. therein) the equations of motion for superbranes in a number of important cases amount to the so-called “geometro-dynamical” constraint which, in the PBGS language, is just a kind of the inverse Higgs constraints. For instance, this applies to the $`N=1,D=10`$ 5-brane . In this case the condition like (9), besides eliminating the Goldstone fermion superfield in terms of the appropriate analog of the $`d=3`$ superfield $`q`$ ($`d=6`$ hypermultiplet superfield), also yields the equation of motion for the latter <sup>4</sup><sup>4</sup>4It is curious that this equation, in accord with the general reasoning of the present work, proved to be finally written in terms of the covariant quantities of nonlinear realization of the pure $`N=1,D=10`$ Poincaré superalgebra, despite the fact that we started in from the coset of the extended supergroup involving $`D=10`$ Lorentz group.. However, as we saw in the above examples, in other interesting cases the inverse Higgs (or geometro-dynamical) constraints do not imply any dynamics which, however, can still be implemented in a manifestly covariant way using the approach proposed here. It still remains to fully understand why in the PBGS scheme the dynamical worldvolume superfield equations are not sensitive to the presence or absence of the automorphism generators in the initial coset construction. This is in contrast with the case of purely bosonic $`p`$-branes. For the self-consistent description of them in terms of nonlinear realizations one should necessarily make use of the cosets of the full target Minkowski space Poincaré group including the Lorentz (automorphism) part of the latter . A possible explanation of this apparent disagreement is that the Goldstone fermion superfields or Goldstone superfields associated with the central charges (and/or with the transverse components of the full momenta) already accommodate the Lorentz and other automorphism groups Goldstone fields. These come out as component fields in the $`\theta `$ -expansion of the Goldstone superfields. So the automorphism groups Goldstone fields are implicitly present in the superbrane superfield equations of motion. The most interesting practical application of the approach exemplified here is the possibility to construct, more or less straightforwardly, the equations for the $`N=4`$ and $`N=8`$ supersymmetric Born-Infeld theory.This work is in progress now . Acknowledgements. This work was supported in part by the Fondo Affari Internazionali Convenzione Particellare INFN-JINR, grants RFBR-CNRS 98-02-22034, RFBR 99-02-18417, INTAS-96-0538, INTAS-96-0308 and NATO Grant PST.CLG 974874.
warning/0003/cond-mat0003095.html
ar5iv
text
# High-frequency hopping conductivity in the quantum Hall effect regime: Acoustical studies ## I Introduction It is well known from numerous low-temperature resistivity measurements that the electronic states of a two-dimensional electron gas (2DEG) whose energies are located between two adjacent Landau levels in a perpendicular magnetic field are localized. Consequently, the conductance is determined by electron hopping between the localized states. The hopping mechanism is temperature-dependent. At temperatures 1-4 K, the conductance $`\sigma `$ is usually determined by nearest-neighbor hopping. In this case its temperature dependence is mainly exponential, $`\sigma _{xx}(T)\mathrm{exp}(E/kT)`$ where $`E`$ is the temperature-independent activation energy, see e. g. Refs. and references therein. (We assume that the 2DEG is located in the $`xy`$ plane, the magnetic field is parallel to the $`𝐳`$-axis, and the electric field is along the $`𝐱`$-axis.) At lower temperatures, $`T1`$ K, the electron hopping distance appears to be greater than the typical distance between nearest neighbors: at such low temperatures it becomes difficult to find a neighbor whose energy is close to the initial one within the accuracy $`kT`$. In this so-called *variable-range-hopping* regime the conductivity $`\sigma _{xx}`$ is also exponentially small, but with an effective activation energy $`E`$ which is temperature-dependent. To clarify the nature of the localized states, we study in this paper the two-dimensional high-frequency (hf) conductance of a 2DEG, $`\sigma _{xx}(\omega )`$, by measuring the velocity and the attenuation of surface acoustic waves (SAW) propagating along the $`x`$-direction nearby the electron layer. Acoustic methods are particularly promising for our task since $`|\sigma _{xx}(\omega )|`$ in the hopping regime may be of the same order of magnitude as the SAW velocity $`V`$. Because the screening of the electric fields produced by the SAW is determined by the ratio $`\sigma _{xx}/V`$, the acoustic properties are sensitive to the variations in $`\sigma _{xx}`$. Furthermore, the attenuation and the velocity of the SAW depend on the *complex* conductance, $$\sigma _{xx}(\omega )\sigma _1(\omega )i\sigma _2(\omega ),$$ and hence both the active, $`\sigma _1`$, and the reactive, $`\sigma _2`$, components can be detected. The active component can be then compared to the static conductivity, $`\sigma _{\text{dc}}`$. A pronounced difference will clearly indicate that the electron states are localized. We compare the experimental results for $`\sigma _1(\omega )`$ and $`\sigma _2(\omega )`$ with existing models for the dielectric response of localized states and extract relevant parameters of the latter. The paper is organized as follows. In Sec. II the experimental setup is presented. Experimental results and their discussion are given in Sec. III. They are summarized in Sec. IV. Details of the derivation of the expressions we use are presented in the Appendix. ## II The experimental setup A sketch of the experimental setup is shown in Fig. 1. A SAW propagating along the surface of a piezoelectric crystal is accompanied by a wave of hf electric field. This electric field penetrates into a 2DEG located in a semiconductor heterostructure mounted on the surface. The field produces electrical currents which, in turn, cause Joule losses. As a result, there are electron-induced contributions both to the SAW attenuation and to its velocity. These effects are governed by the complex frequency-dependent conductivity, $`\sigma _{xx}(\omega )`$, which oscillates as a function of the external magnetic field. Hence, specific oscillations will appear both in the SAW attenuation and its velocity. Under reasonable assumptions, the experimental results allow one to determine $`\sigma _{xx}(\omega )`$ as function of the magnetic field and to analyze its properties. In the present work, the attenuation coefficient $`\mathrm{\Gamma }`$ and the relative velocity change, $`\mathrm{\Delta }V/V`$, are measured as functions of perpendicular magnetic field up to 7 T in the temperature interval 1.5-4.2 K. The Si $`\delta `$-doped GaAs/AlGaAs heterostructure samples with sheet densities $`n=(1.32.7)\times 10^{11}`$ cm<sup>-2</sup> and mobilities $`\mu =(12)\times 10^5`$ cm<sup>2</sup>/V$``$s at T=4.2 K were grown by molecular-beam epitaxy, their structures being shown in Fig. 2. ## III Experimental results and their discussion The expressions relating $`\mathrm{\Gamma }`$ and $`\mathrm{\Delta }V/V`$ to the complex conductance, $`\sigma _1i\sigma _2`$, can be extracted from Refs. and . They read $`\mathrm{\Gamma }\text{(dB/cm)}`$ $`=`$ $`4.34AK^2k\gamma ,`$ (1) $`\mathrm{\Delta }V/V`$ $`=`$ $`(AK^2/2)(\delta v/v).`$ (2) Here $`k=\omega /V`$ is the SAW wave vector, $`K^2`$ is the piezoelectric coupling constant of the substrate (Y-cut LiNbO<sub>3</sub>), and $`A`$ $`=`$ $`8b(k)(\epsilon _1+\epsilon _0)\epsilon _0^2\epsilon _se^{2k(a+d)},`$ (3) $`b(k)`$ $`=`$ $`\left(b_1(k)[b_2(k)b_3(k)]\right)^1,`$ (4) $`b_1(k)`$ $`=`$ $`(\epsilon _1+\epsilon _0)(\epsilon _s+\epsilon _0)`$ (6) $`(\epsilon _1\epsilon _0)(\epsilon _s\epsilon _0)e^{(2ka)},`$ $`b_2(k)`$ $`=`$ $`(\epsilon _1+\epsilon _0)(\epsilon _s+\epsilon _0)`$ (8) $`+(\epsilon _1+\epsilon _0)(\epsilon _s\epsilon _0)e^{(2kd)},`$ $`b_3(k)`$ $`=`$ $`(\epsilon _1\epsilon _0)(\epsilon _s\epsilon _0)e^{(2ka)}+(\epsilon _1\epsilon _0)`$ (10) $`\times (\epsilon _s+\epsilon _0)e^{[2k(a+d)]}.`$ In these equations, $`\epsilon _1`$=50, $`\epsilon _0`$=1 and $`\epsilon _s`$=12 are the dielectric constants of LiNbO<sub>3</sub>, of the vacuum and of the semiconductor, respectively. $`a`$ is the finite vacuum clearance between the sample surface and the LiNbO<sub>3</sub> surface, which can be determined from acoustical measurements. $`d`$ denotes the finite distance between the sample surface and the 2DEG layer. The parameters of our experimental set are $`a=0.5\mu \text{m}`$, $`d=0.59\mu \text{m}`$ for the first sample, and $`0.09\mu \text{m}`$ for the second one. The reduced attenuation $`\gamma `$ and velocity variation $`\mathrm{\Delta }V/V`$ are given by the expressions $`\gamma ={\displaystyle \frac{\mathrm{\Sigma }_1}{\mathrm{\Sigma }_1^2+(1+\mathrm{\Sigma }_2)^2}},\mathrm{\Delta }V/V={\displaystyle \frac{1+\mathrm{\Sigma }_2}{\mathrm{\Sigma }_1^2+(1+\mathrm{\Sigma }_2)^2}};`$ (11) $`\mathrm{\Sigma }_i=(4\pi \sigma _i/\epsilon _sV)t(k),t(k)=[b_2(k)b_3(k)]/2b_1(k).`$ (12) The experimental magnetic field dependence of $`\gamma `$ and $`\mathrm{\Delta }V/V`$ for the sample with carrier density $`n=2.7\times 10^{11}`$ cm<sup>-2</sup> and mobility $`\mu =2\times 10^5`$ cm<sup>2</sup>/V$``$ s are shown in Fig. 3. Similar results have been found previously in GaAs/AlGaAs heterostructures. The real and the imaginary parts of the complex conductance are derived from $`\mathrm{\Gamma }`$ and $`\mathrm{\Delta }V/V`$, using Eqs. (1) and (2). The results obtained for $`T=1.5`$ K and acoustic frequency 30 MHz are shown in Fig. 4. As can be seen, $`\sigma _2`$ practically vanishes near half-integer filling factors, i. e. when the Fermi level is close to any Landau level. In such regions $`\sigma _1(\omega )`$, as has been shown in Ref. , appears to be close to the static conductivity, $`\sigma _{\text{dc}}`$. These facts indicate that the electron states are indeed extended. With a further increase of the magnetic field the Fermi level leaves the Landau band, a metal-dielectric transition takes place, and the electrons become localized in the randomly fluctuating potential of the charged impurities. As the Fermi level departs from the Landau level center, $`\sigma _1(\omega )`$ becomes clearly larger than $`\sigma _{dc}`$, see Fig. 5. Such a behavior can be qualitatively interpreted as absorption by large clusters (“lakes”) disconnected from each other. Inside each cluster the absorption is determined by the value of $`\sigma _{dc}`$. Since the area occupied by the clusters is less than the area occupied by the infinite cluster at the mobility edge, the effective $`\sigma _1(\omega )`$ is less than $`\sigma _{dc}`$ at half-integer $`\nu `$. At the same time, $`\sigma _1(\omega )`$ is greater than $`\sigma _{dc}`$ in the same magnetic field because there is no infinite conducting cluster at the Fermi level. The imaginary part, $`\sigma _2(\omega )`$ *increases* as the Fermi level departs from the Landau level’s center. At magnetic fields corresponding to small integer filling factors, when the Fermi level finds itself in-between the adjacent Landau levels, where $`\sigma _{dc}0`$, $`\sigma _2(\omega )`$ becomes about an order of magnitude larger than $`\sigma _1`$, see Fig. 6. Figure 7 depicts the temperature dependences of $`\sigma _1(\omega )`$ at $`f`$=30 MHz in magnetic fields corresponding to the mid-points of the Hall plateaus. One can see a crossover from a smooth temperature dependence at a strong magnetic field (5.5 T), to a rather steep increase with temperature at weaker fields. Such behavior is compatible with the idea that the conductivity consists of two contributions. The first one is due to the extended states near the adjacent upper Landau level, while the second is coming from the localized states at the Fermi level. The relative occupation of the extended states increases with increasing temperature because of thermally activated processes. Obviously, this effect is dominant at small magnetic fields. We now turn to the region of low temperatures and filling factors close to 2, where hopping between the localized states gives the main contribution to dielectric response. To analyze the experimental results we adopt the so-called two-site approximation, according to which an electron hops between states with close energies localized at two different impurity centers. These states form pair complexes which do not overlap. Therefore, they do not contribute to the static conductivity but are important for the ac response. Being a very simple, the two-site model has been extensively studied, see for a review Refs. and references therein. In the following we will use the 2D version of the theory. Some details of the discussion depend on the assumptions regarding both the density of localized states and the relaxation mechanisms of their population. We therefore re-derive the theoretical results in Appendix A. As is well known , there are two specific contributions to the high-frequency absorption. The first contribution, the so-called resonant, is due to direct absorption of microwave quanta accompanied by inter-level transitions. The second one, the so-called relaxational, or phonon-assisted, is due to phonon-assisted transitions which lead to a lag of the levels populations with respect to the microwave-induced variation in the inter-level spacing. The relative importance of the two mechanisms depends on the frequency $`\omega `$, the temperature $`T`$, as well as on sample parameters. The most important of them is the relaxation rate $`\gamma _0(T)`$ of symmetric pairs with inter-level spacing $`E=kT`$. At $`\omega \sqrt{kT\gamma _0/\mathrm{}}`$ the relaxation contribution to $`\sigma _1(\omega )`$ dominates, and only this one will be taken into account. Following the derivation given in Appendix A we obtain $$\sigma _1=\frac{\pi ^2}{2}\frac{g^2\xi ^3\omega e^4}{\epsilon _s}(_T+_\omega /2)^2.$$ (13) Here $`g`$ is the (constant) single-electron density of states at the Fermi level, $`\xi `$ is the localization length of the electron state, $`_T=\mathrm{ln}J/kT`$, $`J`$ is a typical value of the energy overlap integral which is of the order of the Bohr energy, while $`_\omega =\mathrm{ln}(\gamma _0/\omega )`$. Eq. (13) is valid provided that the logarithmic factors are large. Note that the product $`r_\omega =\xi (_T+_\omega /2)`$ is the distance between the sites forming a hopping pair. Note also that (13) is similar to the result obtained in Ref. , but differs from it by some logarithmic factors and a numerical factor of 1/4. The analysis of $`\sigma _2(\omega )`$ is a bit more complicated because virtual zero-phonon transitions give a comparable contribution. The analysis presented in Appendix A leads to the following expression for the ratio $`\sigma _2(\omega )/\sigma _1(\omega )`$, $$\frac{\sigma _2}{\sigma _1}=\frac{2_\omega (_T^2+_T_\omega /2+_\omega ^2/12)+4c_T^2_c}{\pi (_T^2+_T_\omega +_\omega ^2/4)}.$$ (14) Here $`_c=\mathrm{ln}(\mathrm{}\omega _c/kT)`$, $`\omega _c`$ is the cyclotron frequency, and $`c1`$ is a numerical factor depending on the density of states in the region between the Landau levels, see Appendix A. Using the estimate for $`\gamma _0`$ from Ref. , $$\gamma _0=\frac{4\pi e^2K^2kT}{\epsilon _s\mathrm{}^2V},$$ valid for the piezoelectric relaxation mechanism, as well as other parameters relevant to the present experiment, one concludes that in the hopping regime $`\sigma _2\sigma _1`$. This conclusion agrees with the experimental results obtained for the middles of the Hall plateaus at 5.5 T and 2.7 T and ensures that the conductance mechanism in these regions is indeed hopping. Given an experimental value for $`\sigma _1`$, one can obtain from Eq. (13) the localization length $`\xi `$ provided that the single-electron density of states, $`g`$, is known for given values of the magnetic field. This quantity has been obtained from the temperature dependence measurements of the thermally-activated dc conductivity. It has been shown that for small filling factors the density of states in the plateau regions is finite and almost field-independent, see Fig. 8 Using the density of states versus mobility curve from Ref. , obtained for a sample similar to ours, we estimate the density of states as $`g=2.5\times 10^{24}`$ cm$`{}_{}{}^{2}`$erg<sup>-1</sup>. On the other hand, according to Ref. , the density of states as function of the magnetic field $`H`$ can be expressed by the interpolation formula $$g(H)=\frac{g_0}{1+\sqrt{\mu H}},$$ (15) where $`\mu `$ is the mobility of the 2D-electrons while $`g_0=m/(\pi \mathrm{}^2)`$ is the 2D density of states at $`H=0`$. From Eq. (15) we obtain for $`H=5.5`$ T the density of states $`g=1.7\times 10^{24}`$ cm$`{}_{}{}^{2}`$erg<sup>-1</sup>. Using the first estimate for the density of states one obtains $`\xi =6.5\times 10^6`$ cm, that is about 1.6 times greater than the spacer thickness, $`l_{\text{sp}}=4\times 10^6cm`$. On the other hand, it is the spacer width which characterizes the random potential correlation length in the 2DEG layer. Hence, this fact contradicts our interpretation of experimental results in terms of pure nearest-neighbor pair hopping. To solve the discrepancy, we assume that the high-frequency hopping conductivity of the 2DEG channel is shunted by hopping along the doping Si $`\delta `$-layer. This assumption can be substantiated as follows. Let us suppose that at the middle of the Hall plateau $`\sigma _1^{\nu =2}=4\times 10^7`$ Ohm<sup>-1</sup> and $`\sigma _2^{\nu =2}=2.4\times 10^6`$ Ohm<sup>-1</sup> are entirely determined by the hopping conductivity along the Si $`\delta `$-layer. Such a contribution is only weakly dependent on the magnetic field because the latter is too weak to deform significantly the wave functions of the Si-dopants. Then the contributions to $`\sigma _i`$ associated with the 2DEG channel are just the difference between the experimentally measured $`\sigma _i`$ in a given magnetic field and its value at $`\nu =2`$. We now analyze the dependence of the differences $`F_1\sigma _1\sigma _1^{\nu =2}`$ and $`F_2\sigma _2\sigma _2^{\nu =2}`$ on the filling factor $`\nu `$. The plots of $`\mathrm{lg}F_i`$ versus $`\nu `$ are shown in Fig. 9. Both curves approach straight lines, and consequently can be extrapolated to $`\nu =2`$. Using this extrapolation we have obtained $`F_1^{\nu =2}=10^8`$ Ohm<sup>-1</sup> and $`F_2^{\nu =2}=5\times 10^8`$ Ohm<sup>-1</sup>. It should be noticed here that the extrapolated values of $`F_i^{\nu =2}`$ are two orders of magnitude smaller than the values of $`\sigma _i^{\nu =2}`$, associated with the hopping along Si-$`\delta `$-layer. Using the extrapolated values of $`F_1`$ and $`F_2`$ to extract the 2DEG contributions to $`\sigma _1`$ and $`\sigma _2`$, one can calculate the electron localization length at $`\nu =2`$ from Eq. (13). This procedure is corroborated by the fact that the experimental ratio $`F_2/F_1=5`$ is close to the theoretical value 4.2 coming from Eq. (14). The localization length at $`\nu =2`$ obtained in this way is $`\xi =2\times 10^6`$ cm, which is half of the spacer width. This estimate makes realistic the ”two-site model” which we have extensively used. It should be emphasized, however, that from the above value of $`\xi `$ the hopping length $`r_\omega `$ is estimated to be $`1.4\times 10^5`$ cm. Consequently, there is an interplay between hops to the nearest and more remote neighbors. A more rigorous theory for this situation should be worked out. Such a theory should also explain why the magnetic field dependences of $`\sigma _1`$ and $`\sigma _2`$ at the vicinity of $`\nu `$ = 2 appear to be different – the $`\sigma _1(H)`$-dependence is more pronounced than the $`\sigma _2(H)`$-one. According to the two-site model, both are determined by the respective dependence of the localization length on the magnetic field and should be similar. Indeed, their ratio, from Eq. (14), is almost field-independent. It follows from the experimental data that there exists an additional mechanism leading to the pronounced decrease of $`\sigma _2`$ as the Fermi level falls into the extended states region. A probable mechanism is thermal activation of electrons from the Fermi level to the upper Landau band, leading, firstly, to a decrease of the number of pairs responsible for the hopping conductivity, and, secondly to a screening of the electric field amplitude produced by the SAW. We hope to work out a proper quantitative theory in future. ## IV Conclusions The above analysis leads to the following conclusions: * At the vicinity of the Hall plateau centers, high-frequency hopping conductance in the 2DEG layer can be effectively shunted by hopping inside the doping $`\delta `$-layer. * When the shunting effect is properly subtracted, the results appear to be compatible with the nearest neighbor two-site model of hopping conductivity. * The localization length determined at different magnetic fields (and, consequently, different filling factors) by the above method scales as the magnetic length $`a_H=(\mathrm{}/eH)^{1/2}`$. This agrees with the concept of nearest-neighbor hopping. * The interpretation of the imaginary part of the conductivity, $`\sigma _2(\omega )`$, appears more complicated. While the magnetic field dependence of the real part of the hf hopping conductivity of 2D electrons seems to be determined by the magnetic field dependence of the localization length – the slope of $`\mathrm{lg}\xi (H)`$, calculated from the values of $`F_1^{\nu =2}(H)`$ using Eq. (13), is close to the slope of $`\mathrm{lg}\xi (H)`$ in Ref. – the magnetic field dependence of the imaginary part of the $`hf`$ conductivity has been explained so far only qualitatively. A more detailed quantitative analysis, which would include a proper account of the screening of the SAW-induced hf electrical field by both layers, is required. It is worth emphasizing that the acoustic method used in the present work allows the determination of the localization length near the Hall plateau centers.This is very difficult to achieve using a dc technique. ## V Acknowledgments The work is supported by RFFI N 98-02-18280, MNTRF N 97-1043 grants and I. L. Drichko was supported by the grant of Research Council of Norway. We are grateful to Ora Entin-Wohlman for reading the manuscript and useful remarks. ## A Derivation of $`\sigma (\omega )`$ The derivations of the complex $`\sigma (\omega )`$ within the two-site approximation have been extensively discussed, see e. g. Refs. and . However, the resulting formulae differ in some details. These differences are mainly due to different assumptions about the relaxation of the occupation numbers of the localized states. We therefore present here a unified derivation, in order to clarify the various assumptions and notations. Let us characterize the sites by the single-electron energies $`\phi _{1,2}`$ which would be the actual energies if the Coulomb correlation between the occupation numbers is ignored. The two electron energies, in the absence of quantum hybridization of the states, can be specified by 4 terms, $$W_0=0,W_1=\phi _1,W_2=\phi _2,W_3=\phi _1+\phi _2+\frac{e^2}{\epsilon _sr}.$$ Here $`r`$ is the distance between the sites. As shown in Ref. , when the frequency $`\omega `$ and the temperature $`T`$ are low enough, such that $`\mathrm{}\omega ,kTe^2/\epsilon _sr`$ only the two terms $`W_1`$ and $`W_2`$ can be occupied, and we face a situation of a two-level electronic system (TLS). Quantum tunneling hybridizes the two levels, so that the resulting energies are $$W_\pm =\frac{\phi _1+\phi _2}{2}\pm \frac{E}{2},E=\sqrt{\mathrm{\Delta }^2+\mathrm{\Lambda }^2(r)}.$$ Here $`\mathrm{\Delta }=\phi _1\phi _2`$, $`\mathrm{\Lambda }(r)`$ is the energy overlap integral which decays as $`r`$ increases. The effective Hamiltonian corresponding to this situation is $$_{LR}=\frac{1}{2}\left(\begin{array}{cc}\mathrm{\Delta }& \mathrm{\Lambda }\\ \mathrm{\Lambda }& \mathrm{\Delta }\end{array}\right)=\frac{1}{2}\left(\mathrm{\Delta }\sigma _z\mathrm{\Lambda }\sigma _x\right).$$ (A1) Here $`\sigma _i`$ are the Pauli matrices. Diagonalizing this Hamiltonian we obtain $`_0=(E/2)\sigma _z`$. The quantities $`\phi _i`$ are random, and their distributions are specified as follows. The “center-of-gravity”, $`(\phi _1+\phi _2)/2`$ is assumed to be uniformly distributed within a band of width $`e^2/\epsilon _sr`$; and the difference, $`\mathrm{\Delta }\phi _1\phi _2`$, is also assumed to be uniformly distributed within a band much wider than $`kT`$. Since $`d^2r=rdrd\varphi `$ where $`\varphi `$ is the polar angle in the 2DEG plane, we obtain an $`r`$-independent pair distribution function $`𝒫(\mathrm{\Delta },r,\varphi )=g^2e^2/\epsilon _s`$ where $`g`$ is the (constant) single-electron density of states. It is convenient to change the variables from $`\mathrm{\Delta },r,\varphi `$ to $`\mathrm{\Delta },\mathrm{\Lambda },\varphi `$, $$𝒫(\mathrm{\Delta },\mathrm{\Lambda },\varphi )=g^2(e^2/\epsilon _s)|dr_\mathrm{\Lambda }/d\mathrm{\Lambda }|,$$ (A2) where $`r_\mathrm{\Lambda }`$ is the solution of the equation $`\mathrm{\Lambda }(r)=\mathrm{\Lambda }`$. In the presence of an external ac electric field $`𝐄`$ the pair acquires a dipole moment $`\widehat{𝐝}=e\widehat{𝐫}`$, described by the interaction Hamiltonian $`_i=(𝐄\widehat{𝐝})=e(𝐄\widehat{𝐫})`$. This interaction is added to $`\mathrm{\Delta }`$ in the Hamiltonian (A1). In the representation where $`_0`$ is diagonal, the interaction Hamiltonian becomes $$_{\text{int}}=e(𝐄\widehat{𝐫})\left(\frac{\mathrm{\Delta }}{E}\sigma _z\frac{\mathrm{\Lambda }}{E}\sigma _x\right).$$ (A3) The contribution of a pair to the complex $`\sigma (\omega )`$ can be expressed in terms of the complex susceptibility, $`\chi (\omega )=\sigma (\omega )/i\omega `$ which in turn is given by (cf. Ref. ) $`\chi (\omega )`$ $`=`$ $`{\displaystyle \frac{\pi e^4g^2}{\epsilon _s}}{\displaystyle 𝑑\mathrm{\Delta }𝑑\mathrm{\Lambda }r_\mathrm{\Lambda }^2\left|dr_\mathrm{\Lambda }/d\mathrm{\Lambda }\right|}`$ (A5) $`\times \left[\left(\mathrm{\Delta }/E\right)^2\chi _{zz}(\omega )+\left(\mathrm{\Lambda }/E\right)^2\chi _{xx}(\omega )\right].`$ The partial susceptibilities $`\chi _{pq}`$ are given by (cf. Ref. ) $`\chi _{zz}`$ $`=`$ $`{\displaystyle \frac{1}{kT\mathrm{cosh}^2(E/2kT)}}{\displaystyle \frac{i\gamma _{}}{\omega +i\gamma _{}}}`$ (A6) $`\chi _{xx}`$ $`=`$ $`\mathrm{tanh}\left({\displaystyle \frac{E}{2kT}}\right){\displaystyle \underset{\pm }{}}{\displaystyle \frac{\mathrm{}^1}{\omega E/\mathrm{}+i\gamma }},`$ (A7) where $`\gamma `$ and $`\gamma _{}`$ are the proper relaxation rates. $`\chi _{zz}`$ is responsible for the relaxational contribution, while $`\chi _{xx}`$ is responsible for the resonant one. To continue the calculations one needs to specify the spatial dependence of the overlap integral. Let us assume that $`\mathrm{\Lambda }(r)=Je^{r/\xi }`$, where $`\xi `$ is the localization length. Then, $`r_\mathrm{\Lambda }=\xi \mathrm{ln}(J/\mathrm{\Lambda }),|dr_\mathrm{\Lambda }/d\mathrm{\Lambda }|=\xi /\mathrm{\Lambda }`$. At the next step, it is convenient to transform the variables from $`\mathrm{\Delta },\mathrm{\Lambda }`$ to $`E,p=(\mathrm{\Lambda }/E)^2`$, the Jacobian being $`(2p)^1(1p)^{1/2}`$. This results in $`\stackrel{~}{\chi }(\omega ){\displaystyle \frac{\chi (\omega )}{\chi _0}}`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}𝑑ϵ{\displaystyle _0^1}{\displaystyle \frac{dp}{p\sqrt{1p}}}\mathrm{ln}^2\left({\displaystyle \frac{\stackrel{~}{J}}{ϵ\sqrt{p}}}\right)`$ (A9) $`\times \left[p\chi _{xx}(\omega )+(1p)\chi _{zz}(\omega )\right]`$ where $`\chi _0=\pi e^4g^2\xi ^3kT/2\epsilon _s`$, $`ϵ=E/kT`$, and $`\stackrel{~}{J}=J/kT1`$. The following analysis will be based on Eq. (A9). It can be easily shown that under the conditions of the present experiment the only important contribution to the dissipative part of the susceptibility, $`Im\stackrel{~}{\chi }(\omega )`$, is the one coming from the relaxational mechanism, $`\chi _{zz}`$. Thus we have, $`Im\stackrel{~}{\chi }(\omega )`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dϵ}{\mathrm{cosh}^2(ϵ/2)}}{\displaystyle _0^1}{\displaystyle \frac{dp\sqrt{1p}}{p}}`$ (A11) $`\times \mathrm{ln}^2\left({\displaystyle \frac{\stackrel{~}{J}}{ϵ\sqrt{p}}}\right){\displaystyle \frac{\gamma _{}\omega }{\omega ^2+\gamma _{}^2}}.`$ An important feature of the relaxation rate $`\gamma _{}`$ is that $`\gamma _{}(ϵ,p)=p\gamma _0(ϵ)`$, see e. g. Ref. . Since we are interested in the case $`\omega \gamma _0`$, we can put $`ϵ=1,p=p_\omega \sqrt{\gamma _0(1)/\omega }`$ in the argument of the logarithm and take the logarithm out of the integrand. As a result, $$Im\stackrel{~}{\chi }=\pi \left(_T+\frac{1}{2}_\omega \right)^2,$$ (A12) where $`_T=\mathrm{ln}\stackrel{~}{J}1`$ while $`_\omega =\mathrm{ln}(1/p_\omega )1`$. The relaxational contribution to the real part can be written as $`Re\stackrel{~}{\chi }_{zz}(\omega )={\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dϵ}{\mathrm{cosh}^2(ϵ/2)}}{\displaystyle _0^1}{\displaystyle \frac{dp\sqrt{1p}}{p}}\mathrm{ln}^2\left({\displaystyle \frac{\stackrel{~}{J}}{ϵ\sqrt{p}}}\right)`$ (A13) $`\times {\displaystyle \frac{\gamma _{}^2}{\omega ^2+\gamma _{}^2}}2{\displaystyle _{p_\omega }^1}{\displaystyle \frac{dp}{p}}(_T{\displaystyle \frac{1}{2}}\mathrm{ln}p)^2`$ (A14) $`=2_\omega \left(_T^2+{\displaystyle \frac{1}{2}}_T_\omega +{\displaystyle \frac{1}{12}}_\omega ^2\right).`$ (A15) The contribution from $`\chi _{xx}`$ to the real part of the susceptibility is also important. Putting $`\mathrm{}\omega E`$ we obtain $`Re\stackrel{~}{\chi }_{xx}`$ $`=`$ $`2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dϵ}{ϵ}}\mathrm{tanh}(ϵ/2){\displaystyle _0^1}{\displaystyle \frac{dp}{\sqrt{1p}}}\mathrm{ln}^2\left({\displaystyle \frac{\stackrel{~}{J}}{ϵ\sqrt{p}}}\right)`$ (A16) $``$ $`4_T^2{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dϵ}{ϵ}}\mathrm{tanh}\left({\displaystyle \frac{ϵ}{2}}\right).`$ (A17) The last integral diverges logarithmically at its upper limit. That means that the result is substantially dependent on the total structure of the impurity band. Assuming that (i) $`e^2/\epsilon _s\xi _T\mathrm{}\omega _c`$, and (ii) that we are interested in the situation when the Fermi level is in the middle of the gap, we can replace for a very crude estimate $$g^2_0^{\mathrm{}}\frac{dϵ}{ϵ}\mathrm{tanh}\left(\frac{ϵ}{2}\right)\text{by}_0^{\mathrm{}}\frac{dϵ}{ϵ}g^2(ϵ)\mathrm{tanh}\left(\frac{ϵ}{2}\right).$$ According to certain experimental evidence, the density of localized states within the gap is a weak function of the energy. Then an estimate for the above integral can be written as $`cg^2_c`$ where $`_c=\mathrm{ln}(\mathrm{}\omega _c/kT)`$, while $`c1`$ is a correction factor due to energy dependence of the density of states. As a result, we obtain $$Re\stackrel{~}{\chi }_{xx}4c_T^2_c.$$ (A18) Since $`Re\sigma /Im\sigma =Im\chi /Re\chi `$ we obtain Eq. (14).
warning/0003/gr-qc0003022.html
ar5iv
text
# Untitled Document Gravitational Couplings of Intrinsic Spin Bahram Mashhoon Department of Physics and Astronomy University of Missouri-Columbia Columbia, Missouri 65211, USA Abstract The gravitational couplings of intrinsic spin are briefly reviewed. A consequence of the Dirac equation in the exterior gravitational field of a rotating mass is considered in detail, namely, the difference in the energy of a spin-$`\frac{1}{2}`$ particle polarized vertically up and down near the surface of a rotating body is $`\mathrm{}\mathrm{\Omega }\mathrm{sin}\theta `$. Here $`\theta `$ is the latitude and $`\mathrm{\Omega }=2GJ/(c^2R^3)`$, where $`J`$ and $`R`$ are, respectively, the angular momentum and radius of the body. It seems that this relativistic quantum gravitational effect could be measurable in the foreseeable future. 1 Introduction About forty years ago, Kobzarev and Okun considered the theoretical possibility that a nuclear particle may possess a gravitoelectric dipole moment. This would lead to a violation of the equivalence principle through an interaction of the form $`H_{\mathrm{int}}=A𝝈𝐠`$, where $`A`$ is an amplitude, $`𝝈`$ is the particle spin and g is the gravitational acceleration due to a massive body such as the Earth. Similar spin-gravitoelectric couplings of the form $`f(r)𝝈\widehat{𝐫}`$ have been considered by a number of authors in connection with the possible breakdown of parity and time reversal invariance in gravitation . Leitner and Okubo used the hyperfine splitting of the ground state of hydrogen to set an upper limit on the strength of such an interaction . Meanwhile, Dabbs et al. studied the free fall of neutrons polarized vertically up and down in the gravitational field of the Earth and found no splitting in the gravitational acceleration greater than a few percent of $`g`$. However, a few years later observational evidence was reported for the gravitoelectric dipole moment of the proton . This was soon shown to be spurious by the experiments of Vasil’ev and Young . In particular, Young placed an upper limit of 0.3 Hz on the gravity shift of the proton Larmor frequency in 1969. Finally, a significant upper limit of $`10^4`$ Hz was placed on a possible shift of the deuteron Larmor frequency due to the Earth’s gravitational field by Wineland and Ramsey in 1972. In 1989, the observation of an anomalous difference in the weight of mechanical gyroscopes rotating vertically upward and downward was reported . Again, the existence of such a rotor weight change was soon contradicted by subsequent experiments . The observational search for the role of intrinsic spin in the gravitational interaction as well as the spacetime torsion has continued and many significant experiments have been performed \[10-16\]. These experiments have also explored finite-range axionlike interactions, which could be of the $`\widehat{𝐫}𝝈`$ (“monopole-dipole”) form as well as a linear combination of $`𝝈_A𝝈_B`$ and $`\widehat{𝐫}𝝈_A\widehat{𝐫}𝝈_B`$ (“dipole-dipole”) form, and have placed useful restrictions on the parameters of such interactions. Indeed, the past few decades have witnessed the emergence of extremely precise measurement techniques that make it possible to detect frequency shifts of order $`10^9`$ Hz, an improvement of five orders of magnitude over what was possible three decades ago . The aim of the present paper is to discuss the gravitomagnetic coupling of intrinsic spin due to the fact that according to the standard theory a spinning particle possesses a gravitomagnetic dipole moment. This moment couples to the gravitomagnetic field of a rotating mass (such as the Earth) in complete analogy with the $`𝝁𝐁`$ interaction in electrodynamics. Instead of treating the Dirac equation in the exterior gravitational field of a rotating mass, a heuristic derivation of this general interaction is given in sections 2 and 3 on the basis of the gravitational Larmor theorem. For a spin-$`\frac{1}{2}`$ particle near the surface of the Earth, the effect involves a frequency shift of order $`10^{14}`$ Hz. Section 4 contains a brief discussion of the prospects for the measurement of this relativistic quantum gravitational effect. 2 Inertia of Intrinsic Spin Imagine an observer in a laboratory on the Earth using Earth-based coordinate axes to describe the results of measurements. The particles involved in the experiments on the rotating Earth are waves propagating in inertial spacetime and it is natural to assume that they would keep their polarization aspects fixed in the underlying inertial frame. As measured by the observer, however, such intrinsic spin must “precess” in a sense opposite to the sense of rotation of the Earth. The Hamiltonian associated with such motion would be of the form $`H=𝝈𝛀`$, where $`𝛀`$ is the frequency of rotation of the laboratory frame. The existence of such a Hamiltonian would show that intrinsic spin has rotational inertia. In quantum mechanics, mass and spin characterize the irreducible unitary representations of the inhomogeneous Lorentz group. The inertial properties of mass are well known in classical mechanics through various translational and rotational acceleration effects. In quantum mechanics, the inertial properties of mass have been experimentally investigated by a number of authors . It is therefore interesting to consider the inertial properties of spin . The coupling of intrinsic spin with rotation indicated above may be illustrated by a simple example. Imagine an observer rotating counterclockwise with uniform frequency $`\mathrm{\Omega }`$ about the direction of propagation of a plane linearly polarized monochromatic electromagnetic wave of frequency $`\omega \mathrm{\Omega }`$. For instance, the observer could be in an Earth-based laboratory and $`\mathrm{\Omega }`$ would then be the frequency of the proper rotation of the Earth. We neglect gravitational effects in this section and consider all phenomena in a global inertial frame in Minkowski spacetime. Let the observer move on a circle of radius $`r`$ with speed $`c\beta =r\mathrm{\Omega }`$ in the $`(x,y)`$ plane of the inertial frame and let the electric field of the wave be given by the real part of $$𝐄=E_0\widehat{𝐱}e^{i\omega t+ikz},$$ (1) where $`E_0`$ is a constant amplitude, $`𝐤=k\widehat{𝐳}`$ is the wave vector and $`\omega =ck`$. ¿From the viewpoint of the rotating observer, the direction of linear polarization that is fixed in the inertial frame must drift in a clockwise sense about the direction of propagation, i.e. $$𝐄=E_0\left(\mathrm{cos}\mathrm{\Omega }t\widehat{𝐱}^{}\mathrm{sin}\mathrm{\Omega }t\widehat{𝐲}^{}\right)e^{i\omega t+ikz},$$ (2) where $`\widehat{𝐱}^{}=\widehat{𝐱}\mathrm{cos}\mathrm{\Omega }t+\widehat{𝐲}\mathrm{sin}\mathrm{\Omega }t,\widehat{𝐲}^{}=\widehat{𝐱}\mathrm{sin}\mathrm{\Omega }t+\widehat{𝐲}\mathrm{cos}\mathrm{\Omega }t`$ and $`\widehat{𝐳}^{}=\widehat{𝐳}`$ denote the Cartesian coordinate axes in the rotating frame of the observer. Specifically, the two coordinate systems are related by a simple rotation such that $$\widehat{𝐱}+i\widehat{𝐲}=e^{\pm i\mathrm{\Omega }t}\left(\widehat{𝐱}^{}\pm i\widehat{𝐲}^{}\right).$$ (3) The linearly polarized wave (1) is a coherent superposition of a right circularly polarized (RCP) wave and a left circularly polarized (LCP) wave, i.e. $$𝐄=\frac{1}{2}E_0\left(\widehat{𝐱}+i\widehat{𝐲}\right)e^{i\omega t+ikz}+\frac{1}{2}E_0\left(\widehat{𝐱}i\widehat{𝐲}\right)e^{i\omega t+ikz}.$$ (4) From the viewpoint of the rotating observer, these eigenstates of the radiation field remain invariant, $$𝐄=\frac{1}{2}E_0\left(\widehat{𝐱}^{}+i\widehat{𝐲}^{}\right)e^{i(\omega \mathrm{\Omega })t+ikz}+\frac{1}{2}E_0\left(\widehat{𝐱}^{}i\widehat{𝐲}^{}\right)e^{i(\omega +\mathrm{\Omega })t+ikz},$$ (5) except that the frequency of the RCP component is perceived to be $`\omega \mathrm{\Omega }`$ while that of the LCP wave is perceived to be $`\omega +\mathrm{\Omega }`$ with respect to inertial time $`t`$. The proper time of the observer is, however, $`\tau =t/\gamma `$, where $`\gamma =\left(1\beta ^2\right)^{1/2}`$. Thus we find that the proper frequencies measured by the observer are $$\omega ^{}=\gamma (\omega \mathrm{\Omega }).$$ (6) Here the Lorentz factor accounts for time dilation, which is all that should happen according to the transverse Doppler effect. Instead, we have in (6) the additional “angular Doppler terms” $`\mathrm{\Omega }`$ that have the following physical origin: In an RCP (LCP) wave, the electric and magnetic fields rotate in the positive (negative) sense about the direction of propagation with frequency $`\omega `$. Since the observer rotates in the positive sense with frequency $`\mathrm{\Omega }`$, it perceives the effective frequency of the RCP (LCP) wave to be $`\omega \mathrm{\Omega }(\omega +\mathrm{\Omega })`$. In the JWKB limit, $`\omega \mathrm{}`$ and the “angular Doppler terms” disappear since $`\mathrm{\Omega }/\omega 0`$. Our heuristic treatment ignores certain relativistic corrections that are not essential for the purposes of this discussion. Writing equation (6) in terms of the photon energy as $`E^{}=\gamma (E\mathrm{}\mathrm{\Omega })`$, we see that the deviation from the simple transverse Doppler effect is due to the coupling of the spin of a circularly polarized photon to the rotation of the observer, since a RCP (LCP) photon carries an intrinsic spin of $`\mathrm{}(\mathrm{})`$ along its direction of propagation . These elementary considerations already contain the basic aspects of the phenomenon of spin-rotation coupling, as can be seen from the following discussion based on the theory of relativity . The special theory of relativity consists of two main elements: the principle of relativity (i.e. Lorentz invariance) and the hypothesis of locality. The latter specifies what an accelerated observer measures by establishing a connection between the accelerated observer and an inertial observer. Indeed, it requires that an accelerated observer be at each instant locally equivalent to a momentarily comoving inertial observer. This is a nontrivial axiom since there exist definite acceleration scales of time and length that are associated with an accelerated observer. In the case under consideration, e.g., the acceleration length of the rotating observer is $`=c/\mathrm{\Omega }`$ and the corresponding temporal scale is $`/c=\mathrm{\Omega }^1`$. Moreover, an elementary application of the hypothesis of locality would imply that $`\omega ^{}=\gamma \omega `$ by the transverse Doppler effect, since the connection between the instantaneous inertial frame of the accelerated observer and our global inertial frame simply results in the standard Doppler and aberration formulas with a time-dependent velocity $`c𝜷(t)`$. On the other hand, it should be clear that to measure wave characteristics such as the frequency, one must observe at least a few periods of the oscillations of the wave before a determination of the frequency becomes even possible. In this way, the curvature of the observer’s worldline would have to be taken into consideration, and hence the standard Doppler and aberration formulas of relativity theory are valid only to the extent that the period of the wave $`T=2\pi /\omega `$ is negligible compared to $`\mathrm{\Omega }^1`$, i.e. $`\mathrm{\Omega }T=2\pi \mathrm{\Omega }/\omega 0`$. In view of these remarks, it is therefore natural to apply the locality axiom only to the electromagnetic field; then, the measured field could be Fourier analyzed – which is a nonlocal operation – to obtain the frequency and wave vector content of the field in the accelerated frame. This is indeed the physical basis for the extension of relativistic wave equations to accelerated systems; in fact, this extended hypothesis of locality for wave phenomena is equivalent to the assumption of minimal coupling. Using this approach, one finds that for $`\omega \mathrm{\Omega }`$, the standard Doppler and aberration formulas should be modified to $$\omega ^{}=\gamma (\omega c𝜷𝐤)\gamma \widehat{𝐇}𝛀,$$ (7) $$𝐤^{}=𝐤+(\gamma 1)(\widehat{𝜷}𝐤)\widehat{𝜷}\frac{1}{c}\gamma \omega 𝜷+\frac{1}{c}\gamma (\widehat{𝐇}𝛀)𝜷,$$ (8) where $`\widehat{𝐇}=\pm c𝐤/\omega `$ is the unit helicity vector. One can then consider optical interferometry in a rotating frame that would be based on the spin of the photon in contrast to the Sagnac effect that is connected to its orbital angular momentum . The general expression for spin-rotation coupling can be written as $$E^{}=\gamma (E\mathrm{}M\mathrm{\Omega }),$$ (9) where $`M`$ is the total (orbital plus spin) “magnetic” quantum number along the axis of rotation; that is, $`M=0,\pm 1,\pm 2,\mathrm{}`$ for a scalar or a vector field while $`M\frac{1}{2}=0,\pm 1,\pm 2,\mathrm{}`$ for a Dirac field. In the JWKB approximation, equation (9) can be written as $`E^{}=\gamma (E𝛀𝐉)`$, where $`𝐉=𝐋+𝐒=𝐫\times 𝐏+𝐒`$. Thus $`E^{}=\gamma (E𝐯𝐏)\gamma 𝐒𝛀`$, so that in the absence of intrinsic spin we recover the classical expression for the energy of a particle as measured in the rotating frame with $`𝐯=𝛀\times 𝐫`$. The energy and momentum of a spinning particle as measured by an accelerated observer are then $$E^{}=\gamma (E𝐯𝐏𝐒𝛀),$$ (10) $$𝐏^{}=𝐏+(\gamma 1)(𝐏\widehat{𝜷})\widehat{𝜷}\frac{1}{c}\gamma E𝜷+\frac{1}{c}\gamma (𝐒𝛀)𝜷,$$ (11) using the same JWKB approach as in the derivation of equations (7) - (8). It follows that $$E^2c^2P^2=m^2c^42E(\text{S}𝛀)+(𝐒𝛀)^2.$$ (12) These results reduce to the equations appropriate for light once we set $`E=\mathrm{}\omega ,𝐏=\mathrm{}𝐤,𝐒=\mathrm{}\widehat{𝐇}`$ and $`m=0`$. Experimental evidence for helicity-rotation coupling exists in the microwave and optical regimes via the phenomenon of frequency shift of polarized radiation . Moreover, there is observational evidence for the coupling of spin-$`\frac{1}{2}`$ particles with the rotation of the Earth . The analogous gravitational coupling of intrinsic spin is considered in the next section. 3 Spin-Gravitomagnetic Coupling To extend the physics of spin-rotation coupling to the gravitational field, one must resort to Einstein’s heuristic principle of equivalence. It is possible to interpret this principle in the post-Newtonian approximation via the gravitational Larmor theorem . Newton’s law of gravitation is formally analogous to Coulomb’s law of electricity; therefore, one may describe Newtonian gravitational effects in terms of a gravitoelectric field. The classical tests of general relativity are all due to post-Newtonian gravitoelectric corrections. However, any consistent framework that brings Newtonian gravitation and Lorentz invariance together must of necessity contain a gravitomagnetic field that would be due to mass current. A direct measurement of the gravitomagnetic field of the Earth via the precession of superconducting gyroscopes in a polar orbit about the Earth is one of the goals of the Stanford gyroscope experiment (GP-B) planned for 2001. In the linear approximation of general relativity, where gravitational effects are treated as linear perturbations in a global inertial frame in Minkowski spacetime, one can express the gravitational field equations as Maxwell’s equations for the gravitoelectric field $`𝐄_g`$ and the gravitomagnetic field $`𝐁_g`$ once $`O(c^4)`$ terms are neglected in the post-Newtonian metric perturbations . Specifically, we let $`g_{\mu \nu }=\eta _{\mu \nu }+h_{\mu \nu }`$, where $`\eta _{\mu \nu }`$ is the Minkowski metric and for the linear perturbation $`h_{\mu \nu }`$ we define $`\overline{h}_{\mu \nu }=h_{\mu \nu }\frac{1}{2}\eta _{\mu \nu }h_\alpha ^\alpha `$. Then $`\overline{h}^{00}=4\varphi _g/c^2,\overline{h}^{0i}=2A_g^i/c^2`$ and $`\overline{h}^{ij}=O(c^4)`$. Here $`\varphi _g(t,𝐱)`$ is the gravitoelectric potential and $`𝐀_g(t,𝐱)`$ is the gravitomagnetic vector potential. That is, of the ten effective gravitational potentials $`\overline{h}_{\mu \nu }`$ in general relativity, we neglect the six spatial potentials $`\overline{h}_{ij}`$ as these are of $`O(c^4)`$ for nonrelativistic (astronomical) sources and from the remaining four potentials one can construct a consistent theory of gravitoelectromagnetism (GEM) in this approximation scheme. Let us note that $`\overline{h}^{0\mu }=2c^2(2\varphi _g,𝐀_g)`$, so that the Lorentz gauge condition $`\overline{h}_{,\nu }^{\mu \nu }=0`$ reduces in this case to $$\frac{2}{c}\frac{\varphi _g}{t}+\mathbf{}𝐀_g=0.$$ (13) Thus $`A^\mu =(2\varphi _g,𝐀_g)`$ is the effective GEM potential and the spacetime metric is given by $$ds^2=c^2\left(1\frac{2\varphi _g}{c^2}\right)dt^2\frac{4}{c}\left(𝐀_gd𝐱\right)dt+\left(1+\frac{2\varphi _g}{c^2}\right)\delta _{ij}dx^idx^j.$$ (14) The analogy with electrodynamics turns out to be exact, except for the fact that the ratio of the gravitomagnetic charge to the gravitoelectric charge is two, $`q_B/q_E=2`$; that is, linear gravity is a spin-2 field in contrast to the spin-1 character of the electromagnetic field that implies $`q_B/q_E=1`$ for the Maxwell theory. In electrodynamics, Larmor established a theorem regarding the local equivalence of magnetism and rotation for all charged particles with the same charge-to-mass ratio $`q/m`$. In fact, the electromagnetic field can be locally replaced by an accelerated frame with translational acceleration $`𝐚_L=(q/m)𝐄`$ and rotational (Larmor) frequency $`𝝎_L=q𝐁/(2mc)`$. In electrodynamics, $`q/m`$ can be positive, zero or negative; however, the gravitational charge-to-mass ratio is universal due to the experimentally well-tested equivalence of gravitational and inertial masses. This leads directly to Einstein’s principle of equivalence and hence a geometric theory of gravitation. Einstein’s heuristic principle of equivalence traditionally involves the local equivalence of the gravitoelectric field with the translational acceleration of the “Einstein elevator” in Minkowski spacetime. The interpretation of Einstein’s principle in terms of the gravitational Larmor theorem would then involve, in addition, the local equivalence of the gravitomagnetic field with the Larmor rotation of the elevator as well. Let us consider the exterior field of an almost spherical rotating astronomical body (such as the Earth) with GEM potentials $$\varphi _g\frac{GM}{r},𝐀_g\frac{G}{c}\frac{𝐉\times 𝐫}{r^3},$$ (15) where $`M`$ is the mass and $`J`$ is the angular momentum of the source. These potentials can be obtained from the electromagnetic analogy by assuming that the source has positive gravitoelectric charge $`Q_E=M`$ and gravitomagnetic charge $`Q_B=2M`$. The GEM fields are then $$𝐄_g=\mathbf{}\varphi _g\frac{1}{2c}\frac{}{t}𝐀_g,𝐁_g=\mathbf{}\times 𝐀_g.$$ (16) The motion of test particles in the gravitational field of a rotating mass can be obtained from the Lorentz force law if we assume that for a test particle of inertial mass $`m`$ the gravitational charges are negative, i.e. $`q_E=m`$ and $`q_B=2m`$, in order to take due account of the dominant gravitational attraction between the test particle and the source. It turns out that an ideal test gyroscope at rest outside the rotating source undergoes gravitomagnetic precession $$\frac{d𝐒}{dt}=𝛀_P\times 𝐒$$ (17) with frequency $$𝛀_P=\frac{1}{c}𝐁_g=\frac{GJ}{c^2r^3}\left[3\left(\widehat{𝐫}\widehat{𝐉}\right)\widehat{𝐫}\widehat{𝐉}\right].$$ (18) Imagine now that we replace the gravitomagnetic field by a rotating frame in the neighborhood of the gyroscope. As referred to observers at rest in the rotating frame, the motion of the gyroscope would be the same as before if the observers rotate with Larmor frequency $`𝝎_L=𝛀_P`$. This relation is consistent with the Larmor formula $`𝝎_L=q𝐁/(2mc)`$ once we set $`q_B=2m`$ and $`𝐁_g=c𝛀_P`$ as in equation (18). Thus a consistent and complete gravitoelectromagnetic formalism can be developed along these lines . In particular, the spin-rotation coupling can be extended to gravitomagnetism via the Larmor theorem with $`𝝎_L=𝛀_P`$. That is, the interaction of intrinsic spin with the gravitomagnetic field is given by the Hamiltonian $`H=𝝈𝛀_P`$, since this interaction in the Larmor frame would be $`H=𝝈𝝎_L`$ as described in section 2. Moreover, the Heisenberg equations of motion for the spin-gravity interaction $`H=𝝈𝛀_P`$ are formally the same as equations (17) - (18) for the precession of an ideal test gyroscope. In classical electrodynamics, the magnetic dipole moment for a particle of mass $`m`$ and charge $`q`$ is given by $`𝝁=q𝐒/(2mc)`$, where $`𝐒`$ is its orbital angular momentum. The energy associated with the interaction of this magnetic moment with a magnetic field $`𝐁`$ is $`𝝁𝐁`$. Extending these notions to GEM with $`q_B=2m`$, we find that a gravitomagnetic dipole moment for a gyroscope of spin $`𝐒`$ is $`𝝁_g=𝐒/c`$ and the energy of interaction with a gravitomagnetic field is $`𝝁_g𝐁_g=𝐒𝛀_P`$. A further extension of this result to the intrinsic spin of particles naturally leads to the interaction Hamiltonian $`H=𝝈𝛀_P`$. The gravitoelectric analog of this interaction has already been discussed in section 1; that is, $`H_{\mathrm{int}}=𝐝_g𝐄_g`$, where $`𝐝_g=A𝝈`$ would be the hypothetical gravitoelectric dipole moment of a particle and $`𝐄_g=𝐠`$ from (15) and (16). Let us imagine an experiment in a laboratory near the surface of an astronomical body (such as the Earth) involving the difference in the energy of a particle of spin $`\sigma =s\mathrm{}`$ polarized vertically up and down (i.e. perpendicular to the surface). According to the spin-gravitomagnetic coupling, the result is $$E_+E_{}=2s\mathrm{}\mathrm{\Omega }\mathrm{sin}\theta .$$ (19) Here $`\theta `$ is the geographic latitude (i.e. $`E_+=E_{}`$ at the equator) and $`\mathrm{\Omega }`$ is the effective frequency associated with the gravitomagnetic field $$\mathrm{\Omega }=\frac{2GJ}{c^2R^3},$$ (20) where $`R`$ is the mean radius of the body. Equation (19) expresses a relativistic quantum gravitational effect; indeed, one can write $`\mathrm{}\mathrm{\Omega }=(2cJ/R^3)L_P^2`$, where $`L_P=(\mathrm{}G/c^3)^{1/2}`$ is the Planck length. Let us note that for the Earth $`\mathrm{}\mathrm{\Omega }_E2\times 10^{29}\mathrm{eV}`$, while near the surface of Jupiter $`\mathrm{}\mathrm{\Omega }_J10^{27}\mathrm{eV}`$; similarly, for the Sun $`\mathrm{}\mathrm{\Omega }_S10^{27}\mathrm{eV}`$, but for a neutron star $`\mathrm{}\mathrm{\Omega }_{NS}10^{14}\mathrm{eV}`$. It is important to point out that the spin-rotation-gravity coupling has appeared in the work of many authors who have studied wave equations in accelerated systems and gravitational fields . In particular, the $`𝝈𝛀_P`$ interaction under scrutiny in this work first appeared in the work of de Oliveira and Tiomno . The observation of wave phenomena associated with such couplings was first independently investigated in connection with possible limitations of the general theory of relativity in . Dynamics in electromagnetic fields can be generated by the transformation of the momentum via $`p_\mu p_\mu (q/c)A_\mu `$, where $`A_\mu =(\varphi ,𝐀)`$ is the EM potential. The same holds in the GEM case, except that the analog of $`A_\mu `$ is $`(2\varphi _g,𝐀_g)`$. Let us consider, for instance, the motion of electromagnetic waves in the exterior field of a rotating mass. The effective gravitational charge in this case should be determined based on the fact that a photon of energy $`\mathrm{}\omega `$ in “cyclotron” motion has an effective inertial mass of $`\mathrm{}\omega /c^2`$ and hence the effective GEM charges are $`q_E=\mathrm{}\omega /c^2`$ and $`q_B=2\mathrm{}\omega /c^2`$. The eigenvalue problem in gravitomagnetic fields leads to discreteness properties for the modes reminiscent of the Fock-Darwin-Landau levels in a magnetic field. Imagine, for instance, the motion of electromagnetic waves in a gravitomagnetic field characterized by the magnitude of the effective “cyclotron” frequency $`\mathrm{\Omega }_c2GJ/(c^2r^3)`$. It follows from the explicit solution of Maxwell’s equations in this background that the wave functions are proportional to Hermite polynomials. These polynomials vary over a harmonic characteristic lengthscale $`l_g`$ that is given by $$l_g=\frac{c}{(\omega \mathrm{\Omega }_c)^{1/2}}$$ (21) for an electromagnetic mode of frequency $`\omega `$. If in this equation we set $`\mathrm{}\omega =mc^2`$ and $`\mathrm{\Omega }_c`$ = cyclotron frequency in a magnetic field, we recover the magnetic length that is well known in the discussion of the motion of a charged particle of mass $`m`$ in a magnetic field. It is interesting to note that the gravitomagnetic acceleration length is given by $`_g=c/\mathrm{\Omega }_c`$, so that the gravitomagnetic length (21) is the geometric mean of the reduced wavelength of radiation $`\mathrm{¯}\lambda `$ and $`_g`$. The gravitomagnetic length $`l_g`$ is essentially the same as the radius of the “cyclotron” orbit for a mode with frequency equal to the “cyclotron” frequency ($`\omega =\mathrm{\Omega }_c`$). The eigenvalue spectrum clearly shows the existence of a gravitomagnetic coupling between the photon spin and the rotation of the source . One can show that in the eikonal approximation the gravitational helicity-rotation coupling leads to a differential deflection of polarized radiation thus violating the universality of free fall in a gravitational field beyond the geometric optics limit . That the spin-gravity interaction violates the universality of free fall is already apparent from $`H=𝝈𝛀_P`$, since this Hamiltonian depends only on the spin of the particle and is independent of its mass. Imagine, for instance, the scattering of electromagnetic radiation by a black hole (i.e. pure geometry free of matter). For a Schwarzschild black hole, the scattering amplitude is independent of the polarization of the incident radiation, hence the polarization properties of the radiation are preserved in the scattering process. For a Kerr black hole, however, the scattering amplitude is dependent upon the polarization of the incident radiation. It is possible to give only rough and partial estimates for the motion of wave packets in a gravitomagnetic field . The influence of helicity-rotation coupling on the gravitational deflection of electromagnetic radiation is rather weak and far below the existing observational upper limits , but could become important in future microlensing experiments with polarized radiation. To provide useful astrophysical estimates of the resulting polarization-dependent deflection of radiation, an eikonal approach has been developed for the motion of rays based on equations (7) and (8), i.e. $`\omega (𝐫,𝐤)=ck\pm \widehat{𝐤}𝛀_P(𝐫)`$, so that the Einstein deflection is ignored for the sake of simplicity and only the helicity-rotation coupling is taken into account . In this treatment, the total differential deflection of positive and negative helicity rays approaching the source together from asymptotic infinity and traveling to infinity after deflection vanishes in contrast to what is expected from the wave treatment; however, it is possible to obtain useful estimates for radiation originating near the source. For instance, consider radiation originating over a pole and propagating normal to the rotation axis with an impact parameter $`D`$; then, RCP and LCP waves separate by a total angle of $`\delta 4\mathrm{¯}\lambda GJ/(c^3D^3)`$ about the average Einstein deflection angle. A qualitative description of this effect is given in . The gravitomagnetic splitting $`\delta `$ is small; it amounts to about one milliarcsecond for GHz radio waves passing over a pole of a neutron star. In addition to this splitting, one expects a wavelength-independent gravitomagnetic rotation of the plane of polarization along a ray, i.e. the Skrotskii effect that is the gravitational analog of the Faraday effect . Moreover, the difference in the arrival times of positive and negative helicity radiation originating near a rotating mass and propagating freely outward to a distant point is estimated to be $`T_+T_{}=2\mathrm{¯}\lambda G𝐉𝐫/(c^4r^3)`$, where $`𝐫`$ is the position vector of the point of origin of the radiation relative to the center of the rotating source. This differential time delay due to the different phase speeds of RCP and LCP waves is too small to be measurable at present . The violation of the universality of free fall is a wave effect, so that it vanishes in the $`\mathrm{¯}\lambda /_g0`$ limit. Consider, for instance, a spinning particle in a gravitomagnetic field with the interaction Hamiltonian $`H=𝝈𝛀_P`$. This potential energy is position dependent; therefore, there exists a gravitomagnetic Stern-Gerlach force $`𝐅=\mathbf{}H`$ acting on the particle that is independent of mass and hence violates the universality of the gravitational acceleration. Specifically, $$𝐅=\frac{3GJ}{c^2r^4}\left\{[5(𝝈\widehat{𝐫})(\widehat{𝐉}\widehat{𝐫})𝝈\widehat{𝐉}]\widehat{𝐫}(𝝈\widehat{𝐫})\widehat{𝐉}(\widehat{𝐉}\widehat{𝐫})𝝈\right\},$$ (22) so that the weight operator for the particle $`W=mg𝐅\widehat{𝐫}`$ is given by $`W=mg3H/r`$. If the spin is polarized vertically up or down in a laboratory near the Earth, $$W_\pm =mg\frac{3s}{R}\mathrm{}\mathrm{\Omega }\mathrm{sin}\theta ,$$ (23) so that $`W_\pm =mg(1ϵ`$), where $`ϵ`$ can be expressed as $$ϵ=6s\left(\frac{I}{MR^2}\right)\left(\frac{\mathrm{}\omega }{mc^2}\right)\mathrm{sin}\theta .$$ (24) Here $`J=I\omega `$, $`I`$ is the moment of inertia and $`\omega `$ is the proper rotation frequency of the Earth. For a neutron near the Earth’s surface, $`\mathrm{}\omega /(m_nc^2)5\times 10^{29}`$; hence, $`ϵ`$ is too small to be measurable in the foreseeable future. It follows that for polarized materials the relevant $`ϵ`$ is expected to be even smaller. Let us note that $`ϵ`$ is directly proportional to $`\mathrm{}\omega /(mc^2)`$, which can be expressed as the ratio of the Compton wavelength of the particle ($`\mathrm{}/mc`$) to the rotational acceleration length of the observer ($`c/\omega `$). Indeed, the extended nature of the particle makes it possible for its intrinsic spin to couple to the spacetime curvature resulting in the force $`𝐅`$ that has an exact analog in the classical Mathisson-Papapetrou spin-curvature force . We have thus far discussed the gravitomagnetic spin-rotation coupling in terms of a single rotating source such as the Earth. However, the universality of the gravitational interaction implies that the whole mass-energy content of the universe is involved in every physical experiment via the gravitational interaction. In classical physics, the gravitational force of the rest of the universe enters only through its gradients, which turn out to be rather small for experiments in the solar system. The situation is in general different in quantum physics, however. For instance, in the calculation of the spin-gravity coupling, the gravitomagnetic field generated by the total mass-energy current must be taken into account. This is a difficult problem; however, to get some idea of what is involved here we may use the linear approximation to write the interaction Hamiltonian as $$H=\frac{G}{c^2}\underset{a}{}\frac{3(𝐫_a𝐉_a)(𝐫_a𝝈)r_a^2(𝐉_a𝝈)}{r_a^5},$$ (25) where the sum is over all astronomical sources and $`𝐫_a=𝐱_a𝐱`$ is the vector of relative separation between the particle of spin $`𝝈`$ at $`𝐱`$ and the center of mass of the source $`a`$. Equation (25) can be expressed as $`H=c(𝝈_𝐱)\mathrm{\Phi }_g`$, where $$\mathrm{\Phi }_g=\frac{G}{c^3}\underset{a}{}\frac{𝐉_a𝐫_a}{r_a^3}$$ (26) is the net dimensionless scalar gravitomagnetic potential defined by $`𝐁_g=c^2\mathbf{}\mathrm{\Phi }_g`$. For a laboratory experiment near the Earth, it is simple to show that the net contribution due to the Sun, the Moon and the other planets is negligible. Therefore, to compute $`\mathrm{\Phi }_g`$ one must investigate the cosmic mass-current distribution. This is a difficult observational problem and much remains unknown regarding the distribution of angular momentum in the universe. It is likely that over the largest scales no preferred sense of rotation would be discernible. These considerations lead one to surmise that near the Earth (or Jupiter) the main contribution to the Hamiltonian is simply due to the Earth (or Jupiter), though a completely satisfactory resolution is not available. Conversely, observational data regarding the gravitomagnetic spin-rotation coupling could in principle set limits on the cosmic mass-current distribution. 4 Discussion The gravitational coupling of intrinsic spin with rotation has been described in this work and the consequences of the gravitomagnetic interaction $`H=𝝈𝛀_P`$ have been pointed out. In particular, the gravitomagnetic shift in the Larmor frequency of a nuclear particle has been estimated. Efforts are under way to improve the sensitivity of measurement of such frequency shifts by several orders of magnitude. This could potentially make the effect measurable near the surface of Jupiter . Let us recall that for Jupiter $`\mathrm{}\mathrm{\Omega }_J10^{27}\mathrm{eV}`$, corresponding to a gravitomagnetic Larmor shift of about $`3\times 10^{13}\mathrm{Hz}`$. In view of the current interest in planetary exploration, it appears that the gravitomagnetic coupling of intrinsic spin with rotation could be measurable in the foreseeable future. Acknowledgements I am grateful to Friedrich Hehl and Michael Romalis for helpful discussions and correspondence. References Kobzarev I Yu and Okun LB 1963 JETP 16 1343 Leitner J and Okubo S 1964 Phys. Rev. B 136 1542 Hari Dass ND 1976 Phys. Rev. Lett. 36 393 Peres A 1978 Phys. Rev. D 18 2739 Morgan TA and Peres A 1962 Phys. Rev. Lett. 9 79 Dabbs JWT, Harvey JA, Paya D and Horstmann H 1965 Phys. Rev. B 139 756 Velyukhov GE 1968 JETP Lett. 8 229 Vasil’ev BV 1969 JETP Lett. 9 175 Young BA 1969 Phys. Rev. Lett. 22 1445 Wineland DJ and Ramsey NF 1972 Phys. Rev. A 5 821 Hayasaka H and Takeuchi S 1989 Phys. Rev. Lett. 63 2701 Faller JE, Hollander WJ, Nelson PG and McHugh MP 1990 Phys. Rev. Lett. 64 825 Quinn TJ and Picard A 1990 Nature 343 732 Nitschke JM and Wilmarth PA 1990 Phys. Rev. Lett. 64 2115 Adelberger EG, Heckel BR, Stubbs CW and Rogers WF 1991 Ann. Rev. Nucl. Part. Sci. 41 269 Wineland DJ et al. 1991 Phys. Rev. Lett. 67 1735 Venema BJ et al. 1992 Phys. Rev. Lett. 68 135 Ritter RC, Winkler LI and Gillies GT 1993 Phys. Rev. Lett. 70 701 Berglund CJ et al. 1995 Phys. Rev. Lett. 75 1879; Youdin AN et al. 1996 Phys. Rev. Lett. 77 2170 Ni W-T et al. 1999 Phys. Rev. Lett. 82 2439 Chui TCP and Ni W-T 1993 Phys. Rev. Lett. 71 3247 Vorobyov PV and Gitarts Ya I 1988 Phys. Lett. B 208 146 Bobrakov VF et al. 1991 JETP Lett. 53 294 Vorob’ev PV 1994 JETP Lett. 59 510 Jacobs JP et al. 1995 Phys. Rev. A 52 3521 Werner SA, Staudenmann J-L and Colella R 1979 Phys. Rev. Lett. 42 1103 Rauch H and Werner SA 2000 Neutron Interferometry (Clarendon Press, Oxford) Atwood DK, Horne MA, Shull CG and Arthur J 1984 Phys. Rev. Lett. 52 1673 Bonse U and Wroblewski T 1983 Phys. Rev. Lett. 51 1401 Hasselbach F and Nicklaus M 1993 Phys. Rev. A 48 143 Moorhead GF and Opat GI 1996 Class. Quantum Grav. 13 3129 Schwab K, Bruckner N and Packard RE 1997 Nature 386 585 Gustavson TL, Bouyer P and Kasevich MA 1997 Phys. Rev. Lett. 78 2046 Mashhoon B 1995 Phys. Lett. A 198 9 Mashhoon B, Neutze R, Hannam M and Stedman GE 1998 Phys. Lett. A 249 161 Beth RA 1936 Phys. Rev. 50 115 Mashhoon B 1988 Phys. Rev. Lett. 61 2639; 1992 ibid. 68 3812 Mashhoon B 1989 Phys. Lett. A 139 103 Mashhoon B 1990 Phys. Lett. A 143 176 Mashhoon B 1990 Phys. Lett. A 145 147 Mashhoon B 1993 Phys. Rev. A 47 4498 Mashhoon B 1993 Phys. Lett. A 173 347 Mashhoon B, Gronwald F and Theiss DS 1999 Ann. Physik 8 135 Mashhoon B, Gronwald F and Lichtenegger HIM 2000 in Testing Relativistic Gravity in Space, edited by C. Lämmerzahl, C.W.F. Everitt and F.W. Hehl (Springer-Verlag, Berlin); gr-qc/9912027 de Oliveira CG and Tiomno J 1962 Nuovo Cimento 24 672 Mitskievich NV 1969 Physical Fields in General Relativity Theory, in Russian (Nauka, Moscow) Schmutzer E 1973 Ann. Physik 29 75 Barker BM and O’Connell RF 1975 Phys. Rev. D 12 329 Schmutzer E and Plebański J 1977 Fortschr. Phys. 25 37 Hehl FW and Ni W-T 1990 Phys. Rev. D 42 2045 Cai YQ and Papini G 1991 Phys. Rev. Lett. 66 1259; 1992 ibid. 68 3811 Anandan J 1992 Phys. Rev. Lett. 68 3809 Andretsch J and Lämmerzahl C 1992 Appl. Phys. B 54 351 Silverman MP 1992 Nuovo Cimento D 14 857 Huang J 1994 Ann. Physik 3 53 Soares ID and Tiomno J 1996 Phys. Rev. D 54 2808 Ryder LH 1999 Gen. Rel. Grav. 31 775 Mashhoon B 1974 Nature 250 316 Mashhoon B 1974 Phys. Rev. D 10 1059 Mashhoon B 1975 Phys. Rev. D 11 2679 Damour T and Ruffini R 1974 C.R. Acad. Sci. A 279 971 de Logi WK and Kovacs SJ Jr. 1977 Phys. Rev. D 16 237 Leahy DA 1982 Int. J. Theor. Phys. 21 703 Mashhoon B 1987 Phys. Lett. A 122 299 Futterman JAH, Handler FA and Matzner RA 1988 Scattering from Black Holes (Cambridge University Press, Cambridge) Feng LL and Lu T 1991 Class. Quantum Grav. 8 851 Carini P, Feng LL, Li M and Ruffini R 1992 Phys. Rev. D 46 5407 Harwit M et al. 1974 Nature 249 230 Dennison B, Dickey J and Jauncey D 1976 Nature 263 666 Dennison B et al. 1978 Nature 273 33 Mashhoon B 1999 Gen. Rel. Grav. 31 681 Kopeikin S and Mashhoon B 2000 preprint Romalis M 2000 private communication
warning/0003/hep-th0003223.html
ar5iv
text
# References Recent developments in string dualities have suggested that the known five types of superstring theories in ten dimensions and M-theory in eleven dimensions are just different microscopic descriptions of the same underlying physics. This means that if any of the descriptions can be defined nonperturbatively, it is as good as any other to study the dynamics of the universality class of string/M theories. A celebrated first step towards this end has been made by Ref. , where the so-called Matrix theory has been proposed as a nonperturbative definition of M-theory in the infinite momentum frame. Soon after, the IIB matrix model has been proposed as a nonperturbative definition of type IIB superstring theory. Although there are many evidences that support these conjectures, there is no direct proof that these models reproduce the string/M theories perturbatively<sup>1</sup><sup>1</sup>1See, however, Ref. as an attempt in this direction.. If these conjectures are true, there is a hope to understand all the fundamental questions about the Standard Model, including the space-time dimensionality, gauge group, matter contents, the hierarchy problem, the cosmological constant problem, and so on, in terms of the dynamics of these models. So far, the understanding of the vacuum of these models is quite limited. In this Letter, we attempt to extract information about the dimensionality of the space-time which is dynamically generated in the IIB matrix model. The IIB matrix model is a supersymmetric matrix model obtained formally by taking a zero-volume limit of ten-dimensional SU($`N`$) super Yang-Mills theory. Unlike in field theories, the space-time is treated in this model as a dynamical object, which is represented by ten bosonic $`N\times N`$ hermitian matrices, where $`N`$ should be sent to infinity eventually. Therefore, the model has a potential to explain the dynamical origin of the dimensionality of our space-time, which, in the Standard Model, has to be given as an input parameter. If we are ever going to explain our four-dimensional space-time in this way, the vacuum of the IIB matrix model should be dominated by configurations which have only four-dimensional extent<sup>2</sup><sup>2</sup>2This is similar in spirit to a description of our four-dimensional space-time as a “brane” in a higher-dimensional non-compact space-time, which is proposed as an alternative to a more conventional compactification approach.. This means, in particular, that the ten-dimensional Lorentz invariance of the model should be spontaneously broken to a four-dimensional one.<sup>3</sup><sup>3</sup>3The possibility of a non-trivial vacuum structure in string theory has been considered also in . Such an issue can in principle be addressed by performing Monte Carlo simulation, since the model is completely well defined for arbitrary $`N`$ without any cutoff . Indeed the large-$`N`$ dynamics of the four-dimensional version of the IIB matrix model, which can be obtained by a zero-volume limit of four-dimensional super Yang-Mills theory, has been understood to a considerable extent through Monte Carlo simulation . Comparison of the results with those for the corresponding bosonic theory , revealed that supersymmetry indeed plays an important role in the dynamics of large-$`N`$ reduced models as a nonperturbative definition of string theories. Unfortunately, such a direct approach does not work in the IIB matrix model as it stands, due to the fact that the effective action induced by fermions is generically complex, whereas in the above mentioned four-dimensional version, it is real. In general, when the action of a theory has a non-zero imaginary part, the number of configurations needed to extract any information increases as exponential of the system size. This is the notorious sign problem, which occurs also in many other interesting systems related to particle physics, such as theories with a chiral fermion (as in the present case), theories with a $`\theta `$ vacuum, Chern-Simons theories and theories with a finite baryon number density. We emphasize, however, that the problem is a purely technical one, and indeed its complete solution in some particular class of systems has been obtained recently . In order to search for a way out of this difficulty, let us consider a deformation of the IIB matrix model by introducing an integer parameter $`\nu `$, which couples to the imaginary part of the effective action induced by fermions. The deformed model, which reduces to the IIB matrix model at $`\nu =1`$, is well-defined for arbitrary $`\nu `$. It preserves both gauge invariance and Lorentz invariance, and moreover, it preserves the cluster property, which is an important consequence of the supersymmetry of the original model. At $`\nu =0`$, the sign problem disappears and standard Monte Carlo techniques become applicable. Monte Carlo studies up to $`N=512`$ show, however, that the ten-dimensional Lorentz invariance is not spontaneously broken , which suggests that if Lorentz invariance is spontaneously broken in the original model, the imaginary part of the effective action must play a crucial role. This motivated us to consider the opposite extreme, namely $`\nu =\mathrm{}`$. At $`\nu =\mathrm{}`$, the integration over the bosonic matrices is dominated by the saddle-points of the imaginary part of the effective action. We find that the saddle-points are given by configurations which have only eight-dimensional extent. This implies that the ten-dimensional Lorentz invariance is spontaneously broken at least to an eight-dimensional one. Since we still have to integrate over the saddle-point configurations, the question arises whether the remaining eight-dimensional Lorentz invariance further breaks down, say, to a four-dimensional one. To this end, we study the hessian at configurations which have only $`d`$-dimensional extent ($`d8`$) and find that it is zero of order $`(8d)\{2(N^21)16\}`$. In other words, the imaginary part of the effective action becomes more stationary for smaller $`d`$. This gives a huge enhancement to configurations with smaller $`d`$, which is found to cancel exactly the entropical barrier against having such configurations. Considering the effect of the real part of the effective action, which is studied in Ref. using perturbation theory, it is plausible that the vacuum of the model at $`\nu =\mathrm{}`$ is actually given by $`d<8`$. We also find a lower bound on $`d`$ as $`d>2`$. Further information about the vacuum can be obtained by performing explicitly the integration over the saddle-point configurations by Monte Carlo simulation. Naively, one might consider that the sign problem, which already exists in the original model ($`\nu =1`$), becomes maximally severe at $`\nu =\mathrm{}`$. This is not the case, as we will see. The IIB matrix model is formally a zero-volume limit of ten-dimensional pure $`𝒩=1`$ supersymmetric Yang-Mills theory. The action, therefore, is given by $`S`$ $`=`$ $`S_b+S_f,`$ $`S_b`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{tr}\left([A_\mu ,A_\nu ]^2\right),`$ $`S_f`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{tr}\left(\psi _\alpha (𝒞\mathrm{\Gamma }_\mu )_{\alpha \beta }[A_\mu ,\psi _\beta ]\right).`$ (1) $`A_\mu `$ ($`\mu =1,\mathrm{},10`$) and $`\psi _\alpha `$ ($`\alpha =1,\mathrm{},16`$) are $`N\times N`$ traceless hermitian matrices, which can be expanded in terms of the generators $`t^a`$ of SU($`N`$) as $$(A_\mu )_{ij}=\underset{a=1}{\overset{N^21}{}}A_\mu ^a(t^a)_{ij};(\psi _\alpha )_{ij}=\underset{a=1}{\overset{N^21}{}}\psi _\alpha ^a(t^a)_{ij},$$ (2) where $`A_\mu ^a`$ is a real variable and $`\psi _\alpha ^a`$ is a real Grassmann variable. We have made a Wick rotation in the action (1), so that the metric has Euclidean signature. The $`16\times 16`$ matrices $`\mathrm{\Gamma }_\mu `$ are ten-dimensional gamma matrices after Weyl projection, and the unitary matrix $`𝒞`$ is a charge conjugation matrix satisfying $$𝒞\mathrm{\Gamma }_\mu 𝒞^{}=(\mathrm{\Gamma }_\mu )^{};𝒞^{}=𝒞.$$ (3) The model has a manifest ten-dimensional Lorentz invariance, by which we actually mean an SO(10) invariance. $`A_\mu `$ transforms as a vector and $`\psi _\alpha `$ transforms as a Majorana-Weyl spinor. Pure $`𝒩=1`$ supersymmetric Yang-Mills theory can be also defined in 3D, 4D and 6D, as well as in 10D. Hence, by taking a zero-volume limit of these theories, we can define supersymmetric large-$`N`$ reduced models, which are $`D=3,4,6`$ versions of the IIB matrix model. A nontrivial question then concerns whether the integration over the bosonic matrices is convergent, since the integration domain for hermitian matrices is non-compact. A potential danger of divergence, even for finite $`N`$, exists when the eigenvalues of $`A_\mu `$ become large. This issue has been addressed in Ref. using a one-loop perturbative argument. When all the eigenvalues are well separated from each other, one can expand the matrices $`A_\mu `$ and $`\psi _\alpha `$ around diagonal matrices as $$(A_\mu )_{ij}=x_{i\mu }\delta _{ij}+a_{\mu ij};(\psi _\alpha )_{ij}=\xi _{i\alpha }\delta _{ij}+\phi _{\alpha ij},$$ (4) where $`a_\mu `$ and $`\phi _\alpha `$ are matrices containing only off-diagonal elements. One can then integrate over the off-diagonal elements $`a_\mu `$ and $`\phi _\alpha `$ up to one-loop, by fixing the gauge properly and including the Faddeev-Popov ghosts. The integration over the bosonic off-diagonal elements $`a_\mu `$ and the Faddeev-Popov ghosts, gives a logarithmic attractive potential between all the pairs of $`x_i`$ . This potential, however, is exactly cancelled by the contribution of the fermionic off-diagonal elements $`\phi _\alpha `$ due to supersymmetry. This cancellation is responsible for the cluster property of the model , which is important for the interpretation of the model as a string theory. In order to calculate the effective potential for $`x_{i\mu }`$, one still has to integrate over the fermionic diagonal elements $`\xi _{i\alpha }`$, which is nontrivial. In Ref. , it has been shown that the effective potential for $`x_{i\mu }`$ can be given by a branched-polymer like attractive interaction among $`N`$ points in $`D`$-dimensional space-time represented by $`x_{i\mu }`$ ($`i=1,\mathrm{},N`$). Now counting the power of $`x_i`$ in the partition function, the integration measure for $`x_{i\mu }`$ gives $`D(N1)`$ and the integration over $`\xi _{i\alpha }`$ gives $`\frac{3}{2}p(N1)`$, where $`p=2(D2)`$ is the number of real components of the spinor considered for each $`D`$. Therefore, naively one concludes that when $$D(N1)\frac{3}{2}p(N1)=2(3D)(N1)0,$$ (5) the integration over $`x_{i\mu }`$ is divergent and otherwise convergent. This means that the integral is divergent for $`D=3`$ and convergent for $`D=4,6,10`$, irrespectively of $`N`$. This conclusion is in agreement with an exact result available for $`N=2`$ and a numerical result obtained for $`N=3`$ . For $`D=4`$, Monte Carlo simulations show that the conclusion extends to the large-$`N`$ limit . Therefore, it is conceivable that the above conclusion obtained by the one-loop argument holds in general, although it is not a rigorous proof. Going back to the definition of the model (1), let us first integrate over the fermionic matrices $`\psi _\alpha `$, which induces an effective action for $`A_\mu `$. When fermions are real Grassmann variables as in the present $`D=10`$ case, the fermion integral yields a pfaffian of a $`p(N^21)\times p(N^21)`$ matrix. On the other hand, when fermions are complex Grassmann variables as in the $`D=4,6`$ case, the fermion integral yields a determinant of a $`p(N^21)/2\times p(N^21)/2`$ matrix. For $`D=4`$, the fermion determinant is real positive, which allows a direct Monte Carlo study of the model . For $`D=6,10`$, the fermion determinant ($`D=6`$) or pfaffian ($`D=10`$) is complex in general, since fermions are essentially chiral in these cases. This causes the notorious sign problem. There are some exceptions<sup>4</sup><sup>4</sup>4This is because, when $`(N^21)`$ is smaller than $`D`$, one can always make a $`D`$-dimensional rotation to set $`A_\mu =0`$ for $`\mu =1,\mathrm{},D(N^21)`$, since one can regard $`A_\mu ^a`$ as $`(N^21)`$, $`D`$-dimensional real vectors. The statements in $`D=10`$ for $`N=2`$ and $`N=3`$, for example, follow from (14) and (13), respectively. when $`N`$ is small . For $`D=6`$, the fermion determinant is real positive for $`N=2`$. For $`D=10`$, the pfaffian is real positive for $`N=2`$ and real (but not positive definite) for $`N=3`$. Although we consider only the $`D=10`$ case in what follows, the analysis can be readily applied to the $`D=6`$ case as well. We also restrict ourselves to $`N4`$ so that the pfaffian is complex generically. In $`D=10`$, the fermion integral yields $$Z_f[A]=\text{d}\psi \text{e}^{S_f}=\text{Pf},$$ (6) where $$_{a\alpha ,b\beta }=if_{abc}(𝒞\mathrm{\Gamma }_\mu )_{\alpha \beta }A_\mu ^c$$ (7) is a $`16(N^21)\times 16(N^21)`$ anti-symmetric matrix, regarding each of $`(a\alpha )`$ and $`(b\beta )`$ as a single index. The real totally-antisymmetric tensor $`f_{abc}`$ gives the structure constants of SU($`N`$). In what follows, it proves convenient to work with an explicit representation of the gamma matrices given by $`\mathrm{\Gamma }_1=i\sigma _2\sigma _2\sigma _2\sigma _2;\mathrm{\Gamma }_2=i\sigma _2\sigma _2\mathrm{𝟏}\sigma _1;\mathrm{\Gamma }_3=i\sigma _2\sigma _2\mathrm{𝟏}\sigma _3;`$ (8) $`\mathrm{\Gamma }_4=i\sigma _2\sigma _1\sigma _2\mathrm{𝟏};\mathrm{\Gamma }_5=i\sigma _2\sigma _3\sigma _2\mathrm{𝟏};\mathrm{\Gamma }_6=i\sigma _2\mathrm{𝟏}\sigma _1\sigma _2;`$ $`\mathrm{\Gamma }_7=i\sigma _2\mathrm{𝟏}\sigma _3\sigma _2;\mathrm{\Gamma }_8=i\sigma _1\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏};\mathrm{\Gamma }_9=i\sigma _3\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏};`$ $`\mathrm{\Gamma }_{10}=\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏},`$ for which the charge conjugation matrix $`𝒞`$ becomes a unit matrix. Some important properties of the pfaffian $`\text{Pf}(A)`$ are in order. First of all, it is invariant under a ten-dimensional Lorentz transformation, $`A_\mu ^{}=\mathrm{\Lambda }_{\mu \nu }A_\nu `$, where $`\mathrm{\Lambda }`$ is an SO(10) matrix. It is also invariant under a gauge transformation $`A_\mu ^{}=gA_\nu g^{}`$, where $`g`$ is an SU($`N`$) matrix. Under a parity transformation, $$A_{10}^P=A_{10};A_i^P=A_i(\text{for}i=1,\mathrm{},9),$$ (9) the pfaffian becomes complex conjugate, $$\text{Pf}(A^P)=\{\text{Pf}(A)\}^{},$$ (10) since $`(A^P)=(A)^{}`$. As a consequence, the pfaffian $`\text{Pf}(A)`$ is real, when $`A_{10}=0`$. This also means that $`det=(\text{Pf})^2`$ is real positive, when $`A_{10}=0`$. We next consider the case with $`A_9=A_{10}=0`$. In this case, the pfaffian is actually equal to the fermion determinant of eight-dimensional Weyl fermion which we denote as $`det^{(8)}`$. $`^{(8)}`$ is an $`8(N^21)\times 8(N^21)`$ matrix defined as $`_{a\alpha ,b\beta }^{(8)}=if_{abc}(\stackrel{~}{\mathrm{\Gamma }}_\mu )_{\alpha \beta }A_\mu ^c`$, where the eight-dimensional gamma-matrices $`\stackrel{~}{\mathrm{\Gamma }}_\mu `$ after Weyl projection are given, for example, by $`\stackrel{~}{\mathrm{\Gamma }}_1=i\sigma _2\sigma _2\sigma _2;\stackrel{~}{\mathrm{\Gamma }}_2=i\sigma _2\mathrm{𝟏}\sigma _1;\stackrel{~}{\mathrm{\Gamma }}_3=i\sigma _2\mathrm{𝟏}\sigma _3;\stackrel{~}{\mathrm{\Gamma }}_4=i\sigma _1\sigma _2\mathrm{𝟏};`$ (11) $`\stackrel{~}{\mathrm{\Gamma }}_5=i\sigma _3\sigma _2\mathrm{𝟏};\stackrel{~}{\mathrm{\Gamma }}_6=i\mathbf{\hspace{0.17em}1}\sigma _1\sigma _2;\stackrel{~}{\mathrm{\Gamma }}_7=i\mathbf{\hspace{0.17em}1}\sigma _3\sigma _2;\stackrel{~}{\mathrm{\Gamma }}_8=\mathrm{𝟏}\mathrm{𝟏}\mathrm{𝟏}.`$ With this representation, one finds that $`^{(8)}`$ is a real matrix, which means that $`det^{(8)}`$ is real. However, it is not positive definite, as we have checked numerically. In fact, this is the case even for $`A_8=0`$. However, if we take $`A_7=A_8=0`$, it becomes positive definite, since in this case, we have $`det^{(8)}=|det^{(6)}|^2`$, where $`det^{(6)}`$ is the fermion determinant of six-dimensional Weyl fermion . Finally, when $`A_3=A_4=\mathrm{}=A_{10}=0`$, one finds that $`\text{Pf}(A)=0`$ . This can be proved in the following way. First we note that $`(\text{Pf})^2=det=|detU|^{16}`$, where $`U`$ is an $`(N^21)\times (N^21)`$ matrix defined as $`U_{ab}=f_{abc}X^c`$, where $`X^c=A_1^c+iA_2^c`$. Since $`U_{ab}X^b=0`$, the matrix $`U`$ has a zero-eigenvalue, and therefore $`detU=0`$. This completes the proof. Using Lorentz invariance of the pfaffian $`\text{Pf}(A)`$, we summarize some of its important properties as follows. Let us first define sets of “degenerate” configurations $`\mathrm{\Omega }_d`$ as $$\mathrm{\Omega }_d=\{\{A_\mu \};n_\mu ^{(i)}A_\mu =0\text{for}n_\mu ^{(i)}(i=1,\mathrm{},10d)\text{linearly independent}\}.$$ (12) Obviously, $`\mathrm{\Omega }_1\mathrm{\Omega }_2\mathrm{}\mathrm{\Omega }_9`$. Then the statements are $`\text{(a) When}\{A_\mu \}\mathrm{\Omega }_9,\text{Pf}(A)\text{}.`$ (13) $`\text{(b) When}\{A_\mu \}\mathrm{\Omega }_6,\text{Pf}(A)0.`$ (14) $`\text{(c) When}\{A_\mu \}\mathrm{\Omega }_2,\text{Pf}(A)=0.`$ (15) Here we recall that in the IIB matrix model, the space-time is treated as a dynamical object represented by the bosonic matrices $`A_\mu `$. In this regard, generic configurations in $`\mathrm{\Omega }_d`$ describe $`d`$-dimensional space-time. It is very suggestive that the phase of the pfaffian is sensitive to the dimensionality $`d`$ of the space-time. Indeed, the above results will play a crucial role in the following analysis. Let us write the effective action induced by fermion integral as $`\mathrm{\Gamma }_{\text{eff}}=\mathrm{ln}(\text{Pf})`$, which we decompose into the real part $`\mathrm{\Gamma }^{(r)}`$ and the imaginary part $`\mathrm{\Gamma }^{(i)}`$ as $`\mathrm{\Gamma }_{\text{eff}}=\mathrm{\Gamma }^{(r)}+i\mathrm{\Gamma }^{(i)}`$. Then we generalize the theory by introducing an integer parameter $`\nu `$, which couples only to $`\mathrm{\Gamma }^{(i)}`$, the imaginary part of the effective action, as $$Z_\nu =\text{d}A\text{e}^{S_b[A]\mathrm{\Gamma }^{(r)}[A]}\text{e}^{i\nu \mathrm{\Gamma }^{(i)}[A]}.$$ (16) The parameter $`\nu `$ has to be an integer since $`\mathrm{\Gamma }^{(i)}`$ is defined only up to modulo $`2\pi `$. For $`\nu =1`$, the model reduces to the original model. Since $`\mathrm{\Gamma }^{(i)}`$ flips its sign under a parity transformation due to (10), the models with $`\nu `$ and $`\nu `$ are nothing but parity partners. Since $`\mathrm{\Gamma }^{(r)}`$ and $`\mathrm{\Gamma }^{(i)}`$ are both Lorentz invariant and gauge invariant separately, the generalized model is also both Lorentz invariant and gauge invariant. According to the one-loop perturbative argument given below (4), the deformed IIB matrix model continues to be well-defined for arbitrary $`\nu `$, and moreover, it preserves the cluster property, which is an important consequence of the supersymmetry of the original model. Another important point about the generalization we consider is that $`S_b`$ can be calculated exactly and the result is independent of the parameter $`\nu `$. For this, we recall that in the original supersymmetric large-$`N`$ reduced models ($`D=4,6,10`$), one obtains $$S_b=\frac{1}{4}\left(D+\frac{p}{2}\right)(N^21)=\frac{1}{2}(D1)(N^21),$$ (17) using a scaling argument similar to . Due to the fact that $`\mathrm{\Gamma }^{(i)}[A]`$ is invariant under a scale transformation of $`A_\mu `$, one can easily find that the above result is unaltered by the generalization to arbitrary $`\nu `$ for $`D=6,10`$. In particular, this holds true even in the $`\nu \mathrm{}`$ limit. Note, however, that this does not necessarily imply that all the vacuum expectation values of the generalized model are independent of $`\nu `$. Even their convergence in the $`\nu \mathrm{}`$ limit is nontrivial. This point shall be clarified later. We prove a property of $`\mathrm{\Gamma }^{(i)}`$, which turns out to be essential in the analysis of the $`\nu \mathrm{}`$ limit. We first recall that the pfaffian $`\text{Pf}(A)`$ is a polynomial of $`A_\mu ^a`$ of order $`8(N^21)`$. Hence, $`\mathrm{\Gamma }^{(i)}`$ is infinitely differentiable at a configuration, for which the matrix $``$ is invertible. When the configuration belongs to $`\mathrm{\Omega }_d`$ ($`d=1,\mathrm{},8`$), we find that $$\frac{^n\mathrm{\Gamma }^{(i)}}{A_{\mu _1}^{a_1}A_{\mu _2}^{a_2}\mathrm{}A_{\mu _n}^{a_n}}=\frac{1}{2}\frac{^n}{A_{\mu _1}^{a_1}A_{\mu _2}^{a_2}\mathrm{}A_{\mu _n}^{a_n}}\text{Im}\mathrm{ln}det=0\text{for}n=1,\mathrm{},(9d).$$ (18) This is because, up to $`(9d)`$-th order of perturbations, the configuration stays within $`\mathrm{\Omega }_9`$, and therefore $`det`$ remains to be real positive due to (13). In the $`\nu \mathrm{}`$ limit, the integration over $`A_\mu `$ is dominated by the configurations which satisfy the saddle-point equation given by $$0=\frac{\mathrm{\Gamma }^{(i)}}{A_\mu ^a}=\frac{1}{2}\frac{}{A_\mu ^a}\text{Im}\mathrm{ln}det=\frac{1}{2}\text{Im}\text{Tr }\left(^1^{(\mu a)}\right),$$ (19) where $`^{(\mu a)}`$ is an $`A`$-independent $`16(N^21)\times 16(N^21)`$ matrix defined by $$\left(^{(\mu a)}\right)_{b\beta ,c\gamma }=\frac{_{b\beta ,c\gamma }}{A_\mu ^a}=if_{abc}(𝒞\mathrm{\Gamma }_\mu )_{\beta \gamma }.$$ (20) Since the equation (19) gives $`10(N^21)`$ constraints among $`10(N^21)`$ real variables, naively one would expect that only a very small portion of the configuration space survives. This is not the case, however. In fact, we find that all the configurations that belong to $`\mathrm{\Omega }_8`$ satisfy the saddle-point equation (19) due to (18). The existence of solutions to (19) other than of this type, cannot be excluded. It is reasonable, however, to consider that the configurations in $`\mathrm{\Omega }_8`$ dominate in the sense of Lebesgue measure. Hence, at $`\nu =\mathrm{}`$ the full ten-dimensional Lorentz invariance is broken down at least to an eight-dimensional one. In order to examine whether the remaining eight-dimensional Lorentz invariance further breaks down, we still have to integrate over the saddle-point configurations. An important point here is that the imaginary part of the effective action $`\mathrm{\Gamma }^{(i)}`$ for the saddle-point configurations takes only 0 or $`\pi `$ due to (13). This means that there are actually two sequences of $`\nu `$, (I) $`\nu =0,2,4,\mathrm{},2\mathrm{}`$ and (II) $`\nu =1,3,5,\mathrm{},(2\mathrm{}+1)`$, which give two a priori different limiting theories. In order to formulate the integration over the saddle-point configurations, we first rotate them so that they satisfy $`A_9=A_{10}=0`$ and then integrate over $`A_1,A_2,\mathrm{},A_8`$. We define the Hesse matrix for $`\mathrm{\Gamma }^{(i)}[A]`$ within the directions transverse to the integration domain as $$H_{ja,kb}=\frac{^2\mathrm{\Gamma }^{(i)}}{A_j^aA_k^b}=\frac{1}{2}\frac{^2}{A_j^aA_k^b}\text{Im}\mathrm{ln}det=\frac{1}{2}\text{Im}\text{Tr }\left(^1^{(ja)}^1^{(kb)}\right),$$ (21) where $`j,k=9,10`$. We find that the Hesse matrix $`H_{ja,kb}`$ has $`16`$ zero-eigenvalues with eigenvectors corresponding to perturbations $$(\delta A_9^a,\delta A_{10}^a)=(A_1^a,0),(A_2^a,0),\mathrm{},(A_8^a,0),(0,A_1^a),(0,A_2^a),\mathrm{},(0,A_8^a).$$ (22) These zero-modes are a reflection of the fact that the configurations after these perturbations still stay within $`\mathrm{\Omega }_8`$, and thus satisfy the saddle-point equation (19). When $`N`$ is even, the Hesse matrix $`H_{ja,kb}`$ has actually two more zero-eigenvalues. To see this, we first note that $`H_{9a,9b}=H_{10a,10b}=0`$ and $`H_{9a,10b}=H_{10b,9a}`$. Note also that $`H_{9a,10b}=H_{9b,10a}`$ due to (10), which means that an $`(N^21)\times (N^21)`$ matrix $`K_{ab}`$ defined by $`K_{ab}=H_{9a,10b}`$ is antisymmetric. Due to a general property of an antisymmetric matrix of odd size, $`K_{ab}`$ for even $`N`$ should have a zero-eigenvalue, whose corresponding eigenvector we denote as $`\chi ^a`$. Then one finds that $`(\delta A_9^a,\delta A_{10}^a)=(\chi ^a,0),(0,\chi ^a)`$ are eigenvectors of the Hesse matrix $`H_{ja,kb}`$ with zero-eigenvalues. Unlike the 16 zero-modes in (22), these two additional zero-modes, which exist only for even $`N`$, have nothing to do with the symmetry of the space of solutions to the saddle-point equation (19). Hence, they should be considered as accidental zero-modes. Since the saddle-point analysis in such a case becomes more complicated, we restrict ourselves to the odd $`N`$ case in what follows. In order to deal with the 16 zero-modes, which actually correspond to the Lorentz transformation in the $`(i,9)`$ and $`(i,10)`$ planes ($`i=1,2,\mathrm{},8`$), we have to take into account the phase volume analogous to the Faddeev-Popov determinant. In the present case, it can be given by the 16-dimensional volume spanned by the 16 vectors (22) in the configuration space. We denote this phase volume as $`𝒱^{(16)}`$. Thus we arrive at the models, which describe the two limiting theories corresponding to the even/odd $`\nu `$ sequences, $`Z_{\nu =2\mathrm{}}`$ $`=`$ $`{\displaystyle \left(\underset{\mu =1}{\overset{8}{}}\text{d}A_\mu \right)𝒱^{(16)}|J|^{1/2}\text{e}^{S_b}\left|det^{(8)}\right|},`$ (23) $`Z_{\nu =2\mathrm{}+1}`$ $`=`$ $`{\displaystyle \left(\underset{\mu =1}{\overset{8}{}}\text{d}A_\mu \right)𝒱^{(16)}|J|^{1/2}\text{e}^{S_b}det^{(8)}},`$ (24) where $`J`$ denotes the determinant of the Hesse matrix $`H_{ja,kb}`$ after removing the zero-modes. We recall that $`det^{(8)}`$ is real but not positive definite, which makes the two models a priori different. One can see that the exact result (17) for $`S_b`$ can be reproduced from (23) and (24) as it should, by noting that $`𝒱^{(16)}\lambda ^{16}𝒱^{(16)}`$ and $`J\lambda ^{2\{2(N^21)16\}}J`$ under a scale transformation $`A_\mu \lambda A_\mu `$. Let us consider the question whether the remaining eight-dimensional Lorentz invariance of the models (23) and (24) is further broken. For this we show that the hessian $`J`$ becomes zero for configurations in $`\mathrm{\Omega }_7`$ and that the order of zero increases for configurations in $`\mathrm{\Omega }_d`$ with smaller $`d`$. To quantify this statement, let us consider a configuration with $`A_1,\mathrm{},A_d`$ being generic and $`A_{d+1},\mathrm{},A_8`$ being of order $`ϵ`$. We first note that each element of $`H_{ja,kb}`$ for such a configuration becomes of order $`ϵ^{(8d)}`$ due to (18). Therefore, the non-zero eigenvalues of $`H_{ja,kb}`$ become of order $`ϵ^{(8d)}`$. We have checked numerically that they are not of order higher than $`ϵ^{(8d)}`$ generically. This means that $`J`$ is of order $`ϵ^{(8d)\{2(N^21)16\}}`$. On the other hand, degenerate configurations with smaller $`d`$ is suppressed by the entropy factor $`ϵ^{(8d)(N^21)}`$ coming from the integration measure. The phase volume factor $`𝒱^{(16)}`$ gives also a suppression of order $`ϵ^{2(8d)}`$. Collecting all the powers of $`ϵ`$, we find a suppression of order $`ϵ^{10(8d)}`$. This means, first of all, that the integrals (23) and (24) are non-singular at the degenerate configurations. Secondly, we note that, as far as the $`N`$-dependent part is concerned, the suppression for smaller $`d`$ coming the entropy factor $`ϵ^{(8d)(N^21)}`$ is exactly cancelled by the enhancement coming from $`|J|^{1/2}`$. This means that there is essentially no entropical barrier against having smaller $`d`$. So far, we have been focusing on the effect of $`\mathrm{\Gamma }^{(i)}`$ on the dynamics of the IIB matrix model. The effect of $`S_b`$ and $`\mathrm{\Gamma }^{(r)}`$, on the other hand, has been studied in Ref. by using the low-energy effective theory obtained by the one-loop perturbation theory (4). There, it was argued that the complicated branched-polymer interaction among the diagonal elements $`x_{i\mu }`$ of the bosonic matrices $`A_\mu `$ might induce a collapse of the distribution of $`x_{i\mu }`$. Monte Carlo simulation of the low-energy effective theory at $`\nu =0`$ shows that the effect of $`S_b`$ and $`\mathrm{\Gamma }^{(r)}`$ is not sufficient to induce a spontaneous breakdown of Lorentz invariance. However, after taking into account the effect of $`\mathrm{\Gamma }^{(i)}`$ by sending $`\nu `$ to infinity, it is very plausible that the remaining eight-dimensional Lorentz invariance of the models (23) and (24) is further broken by the effect of $`S_b`$ and $`\mathrm{\Gamma }^{(r)}`$, because there is no entropical barrier against having degenerate configurations any more. If this is the case, the vacuum at $`\nu =\mathrm{}`$ is given by degenerate configurations with $`d<8`$. In this regard, we recall also that the pfaffian $`\text{Pf}(A)`$ is zero when $`\{A_\mu \}\mathrm{\Omega }_2`$, as stated in (15). Therefore, the dimensionality $`d`$ of the vacuum configurations must be $`d>2`$. In order to check the above statements and to determine the dimensionality $`d`$ of the vacuum configurations at $`\nu =\mathrm{}`$, one has to carry out the integration over the saddle-point configurations described by (23) and (24), for example, by Monte Carlo simulation. Note that the model (23) is not plagued by the sign problem any more. The model (24), on the other hand, can be studied using the configurations generated with the model (23) as $$𝒪_{\nu =2\mathrm{}+1}=\frac{𝒪\text{sgn}(det^{(8)})_{\nu =2\mathrm{}}}{\text{sgn}(det^{(8)})_{\nu =2\mathrm{}}}.$$ (25) If the fermion determinant $`det^{(8)}`$ does not have a definite sign for dominant configurations, the denominator as well as the numerator of the r.h.s. in (25) becomes very small, exhibiting the sign problem. However, if it turns out that the vacuum of the model (23) is given by configurations with $`d6`$, then due to (14), we have $`\text{sgn}(det^{(8)})=1`$ for dominant configurations. In this case, the sign problem, which a priori exists in the model (24), is solved in a sense dynamically, and the model (24) is actually equivalent to the model (23), since $`𝒪_{\nu =2\mathrm{}+1}=𝒪_{\nu =2\mathrm{}}`$ for any observables $`𝒪`$. To summarize, we considered a deformation of the IIB matrix model by introducing an integer parameter $`\nu `$ which couples to $`\mathrm{\Gamma }^{(i)}`$, the imaginary part of the effective action induced by fermions. We studied the deformed model at $`\nu =\mathrm{}`$, where the integration over the bosonic matrices is dominated by the configurations for which $`\mathrm{\Gamma }^{(i)}`$ is stationary. First of all, there is still a huge configuration space left as the saddle-point configurations which we have to integrate over. Secondly, these saddle-point configurations have more than two shrunken directions. Thirdly, the more shrunken directions the configuration has, the more stationary $`\mathrm{\Gamma }^{(i)}`$ becomes. This gives rise to an enhancement for configurations with more shrunken directions, and the effect was shown to cancel exactly the $`N`$-dependent entropical barrier against having such configurations. An intriguing feature of this enhancement is that it occurs exactly when the configuration becomes a lower-dimensional hyperplane. This may be responsible for generating a flat space-time instead of a curved one or a fractal one, as a result of the dynamics of the IIB matrix model. We argued that the dimensionality $`d`$ of the space-time generated dynamically in the deformed IIB matrix model at $`\nu =\mathrm{}`$ is $`2<d8`$ and most likely $`d<8`$. We derived the models which describe the integration over the saddle-point configurations for the two a priori different limiting theories corresponding to the even/odd $`\nu `$ sequences. Remarkably, the model with the even $`\nu `$ sequence is not plagued by the sign problem. One can therefore study the model by standard Monte Carlo simulation to extract the dimensionality $`d`$ of the vacuum configurations. If this turns out to be $`d6`$, the model with the odd $`\nu `$ sequence is actually equivalent to the model with the even $`\nu `$ sequence. Whether the original model ($`\nu =1`$) belongs to the same phase as $`\nu =\mathrm{}`$ is a nontrivial question, which is not accessible through standard Monte Carlo simulation due to the sign problem. We quote, however, an example from history, where an analogous approach was successful. It is the strong coupling limit in the lattice formulation of nonabelian gauge theories. Although one should send the bare coupling constant to zero in the continuum limit, confinement in these theories has been clearly demonstrated in the strong coupling limit . The limit also provides a qualitative understanding of the spontaneous breakdown of chiral symmetry . The existence of such an approach is indeed one of the advantages of having a nonperturbative formulation, which should also apply to the case of string theory. Note, however, that the above successes in the case of nonabelian gauge theories rely crucially on the fact that there is no phase transition between the strong coupling regime and the weak coupling regime . An illustrative example clarifying this point is that confinement holds in the strong coupling limit even for abelian gauge theories, for which there is actually a phase transition to a deconfining phase at an intermediate coupling constant. The fact that the model at $`\nu =\mathrm{}`$ still has a huge configuration space to integrate over, which is quite peculiar to this system, may suggest that it is in fact quite close to the original model ($`\nu =1`$). Although it is hard to justify this statement rigorously, it would be certainly worthwhile to explore further the dynamics of the deformed IIB matrix model in this limit. J. N. would like to thank K.N. Anagnostopoulos, T. Hotta, T. Izubuchi and A. Tsuchiya for collaborations at the earlier stage of this work. We are also grateful to J. Ambjørn, W. Bietenholz, D. Bödeker, F.R. Klinkhamer and N. Obers for valuable comments and discussions. J. N. is supported by the Japan Society for the Promotion of Science as a Postdoctoral Fellow for Research Abroad. The work of G. V. is supported by MURST (Italy) within the project “Fisica Teorica delle Interazioni Fondamentali”.
warning/0003/quant-ph0003047.html
ar5iv
text
# Indistinguishability and nonlocality in Einstein-Podolsky-Rosen experiment ## I Introduction The recent experiments on teleportation demonstrate that instantaneous transportation of information is an experimental fact. The quantum state of a given system may be transported from one location to another without moving through the intervening space. Partial implementations of quantum teleportation over macroscopic distances have been achieved by using optical systems. That suggests, in our opinion, that a revision on the very structure of space-time is demanded. Actually, teleportation is essentially based on Einstein-Podolsky-Rosen (EPR) experiment . We claim that we should revise to notion of space-time at microscopic levels, by taking into account the non-individuality of elementary particles. If two particles are indistinguishable, then we should not ascribe usual space-time coordinates to them, since these coordinates act like labels. In this paper we consider that ‘indistinguishable’ objects are objects that share their properties, while ‘identical objects’ means ‘the very same object’, and not two objects at all. We need to settle this vocabulary, since in physics textbooks it is usual to consider the words ‘indistinguishable’ and ‘identical’ as synonymous. Relativity does not allow signals with velocities greater than the speed of light in the vacuum. EPR suggests that there is an instantaneous communication between two elementary particles. It is considered that the interference produced by two light beams, in a two-slit experiment, is determined by both their mutual coherence and the indistinguishability of the quantum particle paths. For instance, Mandel has proposed a quantitative link between the wave and the particle descriptions by using an adequate decomposition of the density operator. In this paper we propose the use of a set-theoretical framework without identity for the EPR Gedanken experiment, which allows us to show that nonlocal phenomena between entangled particles is a logical consequence from their non-individuality. The set-theory with no identity that we use is quasi-set theory . Quasi-set theory is based on Zermelo-Fraenkel axioms for sets and permits to cope with collections of indistinguishable objects by allowing the presence of two sorts of atoms (Urelemente), termed $`m`$-atoms and $`M`$-atoms. A binary relation of indistinguishability between $`m`$-atoms (denoted by the symbol $``$), is used instead of identity, and it is postulated that $``$ has the properties of an equivalence relation. The predicate of equality cannot be applied to the $`m`$-atoms, since no expression of the form $`x=y`$ is a well-formed formula if $`x`$ or $`y`$ denote $`m`$-atoms. Hence, there is a precise sense in saying that $`m`$-atoms can be indistinguishable without being identical. We have recently proposed (with D. Krause and A. G. Volkov) that quasi-set theory provides a tool for dealing with collections of indistinguishable elementary particles . In we proved (with A. M. S. Santos) that a Maxwell-Boltzmann distribution is possible even in an ensemble of indistinguishable particles. In this paper we show that indistinguishability implies nonlocality as it follows in the next paragraphs. ## II EPRB The original Einstein-Podolsky-Rosen Gedanken experiment dealt with measurements of position and momentum in a two-particle system. Here we appeal to the use of a composite spin $`\frac{1}{2}`$ system started with D. Bohm. This kind of experimental setup we refer to as Einstein-Podolsky-Rosen-Bohm (EPRB) experiment. Our discussion on this topic is essentially based on the classic textbook by J. J. Sakurai . It is well known that the state ket of a two-electron system in a spin-singlet state can be described by: $$\left(\frac{1}{\sqrt{2}}\right)(|𝐳+;𝐳|𝐳;𝐳+),$$ (1) where $`𝐳`$ is an arbitrary quantization direction, and $`|𝐳+;𝐳`$ means that electron $`1`$ is in the spin-up state while electron $`2`$ is in the spin-down state. Something similar may be said about $`|𝐳;𝐳+`$. If the spin component of particle $`1`$ is shown to be, e.g., in the spin-up state, the other particle component is necessarily in the spin-down state. This remarkable correlation has been experimentaly confirmed . Some authors have discussed the possibility of a comunication between entangled particles by means of ‘wormholes’ . In this paper we suggest another topological explanation of this weird phenomenum. ## III Quasi-Metric Spaces Our mathematical framework for describing EPRB is quasi-sets. Quasi-set theory $`𝒬`$ allows the presence of two sorts of atoms (Urelemente), termed $`m`$-atoms and $`M`$-atoms, identified by two unary predicates $`m(x)`$ and $`M(x)`$, respectively. It is important to observe that the term ‘atom’ is in the mathematical sense. Concerning the $`m`$-atoms, a weaker ‘relation of indistinguishability’ (denoted by the symbol $``$), is used instead of identity, and it is postulated that $``$ has the properties of an equivalence relation. The universe of $`𝒬`$ is composed by $`m`$-atoms, $`M`$-atoms and quasi-sets. The sentence ‘$`x`$ is a quasi-set’ is denoted by $`Q(x)`$, where $`Q`$ is a unary predicate. The axiomatics is adapted from that of ZFU (Zermelo-Fraenkel with Urelemente), and when we restrict the theory in not considering $`m`$-atoms, quasi-set theory is essentially equivalent to ZFU, and the corresponding quasi-sets will be termed ‘ZFU-sets’ (similarly, if also the $`m`$-atoms are ruled out, the theory collapses into ZFC). The $`M`$-atoms play the role of the Urelemente in the sense of ZFU. In order to preserve the concept of identity for the ‘well-behaved’ objects, an extensional equality is introduced for those entities which are not $`m`$-atoms on the following grounds: for all $`x`$ and $`y`$, if they are not $`m`$-atoms, then $`x=_Ey`$ iff $`(Q(x)Q(y)(z(zxzy)))(M(x)M(y)xy)`$. We are using standard logical notation: $``$ is the conditional of propositional calculus, $`\neg `$ is negation, $``$ is conjunction, $``$ is disjunction, $``$ is the biconditional, and $``$ and $``$ are, respectively, the universal and the existential quantifiers of predicate calculus. It is possible to prove that $`=_E`$ has all the properties of classical identity and so these properties hold regarding $`M`$-atoms and ‘ZFU-sets’. It is straight to see that we can easily define the binary relations “$`_E`$”, “$`<_E`$”, “$`>_E`$”, “$`_E`$”, and “$`_E`$”, as natural extensions, respectively, of “$``$”, “$`<`$”, “$`>`$”, “$``$”, and “$``$” in the ZFU-set of, e.g., real numbers. According to the weak-pair axiom in quasi-set theory, for all $`x`$ and $`y`$, there exists a quasi-set whose elements are the indistinguishable objects from either $`x`$ or $`y`$. In symbols: $`xyz(Q(z)t(tztxty))`$. Such a quasi-set is denoted by $`[x,y]`$ and, when $`xy`$, we have $`[x]`$ by definition. We remark that this quasi-set cannot be regarded as the ‘singleton’ of $`x`$, since its elements are all the objects indistinguishable from $`x`$, so its ‘cardinality’ (or quasi-cardinality, if we use the original terminology in ) may be greater than $`1`$. A concept of strong singleton, which plays an important role in the applications of quasi-set theory, may be defined. We call $`[x]`$ a weak singleton. It is rather intuitive that the concept of function cannot be defined in the standard way, so it is introduced a weaker concept of quasi-function, which maps collections of indistinguishable objects into collections of indistinguishable objects. When there are no $`m`$-atoms involved, the concept is reduced to that of function as usually understood. The concept of relation is like the standard one: a quasi-set $`w`$ is a relation between two quasi-sets $`x`$ and $`y`$ if $`w`$ satisfies the following predicate $`R`$: $`R(w)`$ iff $`Q(w)z(zwuv(uxvyz=_Eu,v))`$. The notion of ordered pair $`u,v`$ is analogous to the usual definition. As usual, if $`x=_Ey`$, we say that $`R`$ is a relation on $`x`$. We denote by $`Dom(R)`$ (the domain of $`R`$) the quasi-set $`[ux:u,vR]`$ and by $`Rang(R)`$ (the range of $`R`$) the quasi-set $`[vy:u,vR]`$. Let $`x`$ and $`y`$ be quasi-sets. We say that $`f`$ is a quasi-function from $`x`$ to $`y`$ if $`f`$ is such that ($`R`$ is the predicate for ‘relation’ defined above): $`R(f)u(uxv(vyu,vf))uu^{}vv^{}(u,vfu^{},v^{}fuu^{}vv^{})`$. As a final remark on this review on quasi-sets we say that there is a unary functional letter in the language of quasi-set theory named as $`qc`$. If $`x`$ is a variable, then $`qc(x)`$ corresponds to the quasi-cardinality of $`x`$, which is a cardinal. Roughly speaking, $`qc(x)`$ corresponds to the number of elements of $`x`$ and it is a natural extension of the usual notion of cardinality of a set. Next we define metric spaces into the context of quasi-set theory. Our main goal is to define a very simple quasi-set-theoretical structure for a space-time of a system of two EPRB-correlated particles. ###### Definition 1 A quasi-metric space is an ordered pair $`X,d`$, such that: 1. $`X`$ is a non-empty quasi-set. 2. $`d:X\times X\mathrm{}`$ is a quasi-function which associates each ordered pair $`x,y`$ to a real number $`d(x,y)`$, which we refer to as the distance between $`x`$ and $`y`$. 3. $`d(x,y)=_E0`$ if and only if $`xy`$. 4. $`d(x,y)>_E0`$ if and only if $`\neg (xy)`$, that is, $`x`$ and $`y`$ are distinguishable ($`\neg `$ is the standard negation connective from mathematical logic). 5. $`d(x,y)=_Ed(y,x)`$. 6. $`d(x,z)_Ed(x,y)+d(y,z)`$ The above definition is a very natural generalization of the usual concept of metric space, which certainly deserves an adequate mathematical study that we do not accomplish here, since that task is out of the scope of this paper. Now we present an example of a non-trivial quasi-metric space with its correspondent physical interpretation in terms of the EPRB Gedanken experiment. ###### Definition 2 An EPRB-Space is a quasi-metric space $`𝒮=_EM,d_q`$, where: $`M=_EV[x]_2`$, where $`V\mathrm{}^n`$ is an open set of the usual set of ordered $`n`$-tuples of real numbers, endowed with the euclidian metric;<sup>1</sup><sup>1</sup>1It could be a flat metric of Lorentz signature. $`[x]_2`$ is a weak singleton such that $`m(x)`$; and $`qc([x]_2)=_E2`$. If $`a`$ and $`b`$ are elements of $`V`$ then $`d_E(a,b)_E2c`$, where $`d_E`$ is the euclidian distance, and $`c`$ is a real constant such that $`c>_E0`$. $`d_q(x,a)=_Ed_q(a,x)=_Ec`$ for all $`x[x]_2`$, and for all $`aV`$. $`d_q(a,b)=_Ed_E(a,b)`$ if $`a`$ and $`b`$ are both elements of $`V`$. We know that the entanglement state of the two correlated particles in EPRB entails a relation of indistinguishability between them. So, such particles cannot be labeled by anything. The EPRB-correlated particles cannot be labeled by their coordinates in space-time, since they are indiscernibles and space-time is a classical structure with individual points. In other words, we propose another structure for space-time in quantum mechanics. We suggest that entangled quantum particles define another space-time structure (quasi-metric space $`𝒮`$), quite different from that one in classical mechanics. The space or space-time coordinates in $`V`$ correspond to the coordinates of macroscopic and distinguishable particles or devices. The quasi-set $`[x]_2`$ in axiom A1 of an EPRB space-time corresponds to the ‘coordinates’ (or corresponding quasi-metric space points) associated to the two-particle system of EPRB-correlated electrons. So, any “nonlocal” phenomena among EPRB-correlated particles is not nonlocal at all, since the distance between indiscernibles is always zero, according to axiom A1 in definition (2) and axiom 3 in definition (1). One interesting side result is that our system $`𝒮`$ does not allow us to localize the quantum indistinguishable particles, which are associated to the elements of $`[x]_2`$. That is, entangled particles do not have local coordinates in the usual sense. Nevertheless their distance to any point of $`V`$ is always constant. It is easy to verify that the proper axioms of $`M,d_q`$ are consistent with those of a quasi-metric space. One natural question is: what is the role of constant $`c`$? Our idea is to extend the euclidian metric $`d_E`$ to a space which includes m-atoms $`x`$, i.e., Urelemente $`x`$ such that $`m(x)`$. Since $`d_q(a,x)`$ cannot be zero if $`m(x)`$ and $`a`$ is an ordered $`n`$-tuple of real numbers ($`\neg (xa)`$), then the simplest solution is to consider that $`d_q(a,x)`$ is constant. Nevertheless, if $`d_q(x,a)=_Ec`$, then we cannot consider $`V=_E\mathrm{}^n`$, since in this case we can easily show that assumption $`6`$ in Definition (1) is not satisfied. So, constant $`c`$ is closely related to the open set $`V`$. Since our EPRB-space is restricted to a neighborhood $`V`$ of some points in $`\mathrm{}^n`$, it seems clear that this neighboorhood has something to do with the experimental fact that the spin measurement probe (in EPRB) is localized in a specific point of classical space-time. More specificaly, $`V`$ is the neighborhood of the space-time coordinates of the spin measurement probes. In other words, the region $`V`$ may be regarded as the set-theoretical union of two open balls in $`\mathrm{}^n`$ which are associated to the space-time regions occupied by the two spin measurement devices in the EPRB experiment. See FIG. 1. Such a mathematical model is not in conflict with not-instantaneous electromagnetic interactions between the two correlated electrons if we consider that this null distance between them allows just interactions directly related to the entangled state involved. Besides, it is a nice result that any null distance is obviously invariant under Galilean as well as Lorentzian transformations of coordinates. Despite the fact that simultaneity depends on the observer in special relativity, this simultaneity occurs between two points very close (null distance). It seems natural to admit that a relativistic approach to this kind of space-time structure is demanded. But that is a task for future works. ## IV Conclusions According to FIG. 1 and our previous explanation, nonlocality in EPRB is a misunderstanding. Although the distance between the spin measurement devices is not zero, the distance between the two particles is null. So, there is no non-local phenomena at all. Besides, the distance between a given particle of the EPRB system and any point of $`V`$ (the space-time region occupied by the spin peasurement devices) is always $`c`$. This last result is a confirmation that there is no manner to localize indistinguishable particles in a specific point of space-time, but there is a way to associate these particles to neighborhoods in space-time. Such an association may cause the illusion that the particles in this system can be localized. But they cannot, since EPRB correlated particles are indistinguishable.
warning/0003/hep-ex0003013.html
ar5iv
text
# 1 Introduction ## 1 Introduction Within the framework of the Standard Model of electroweak interactions, the elements of the Cabibbo-Kobayashi-Maskawa mixing matrix are free parameters, constrained only by the requirement that the matrix be unitary. The values of the matrix elements can only be determined by experiment. Heavy Quark Effective Theory (HQET) provides a means to extract the magnitude of the element $`V_{\mathrm{cb}}`$ from particular semileptonic b decays, with relatively small theoretical uncertainties . In this paper, the value of $`|V_{\mathrm{cb}}|`$ is extracted by studying the decay rate for the process $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ as a function of the recoil kinematics of the $`\mathrm{D}^+`$ meson <sup>1</sup><sup>1</sup>1Charge conjugate reactions are always implied, and the symbol $`\mathrm{}`$ refers to either an electron or muon. . The decay rate is parameterised as a function of the variable $`\omega `$, defined as the scalar product of the four-velocities of the $`\mathrm{D}^+`$ and $`\overline{\mathrm{B}}^0`$ mesons. This is related to the square of the four-momentum transfer from the $`\overline{\mathrm{B}}^0`$ to the $`\mathrm{}^{}\overline{\nu }`$ system, $`q^2`$, by $$\omega =\frac{m_{\mathrm{D}^+}^2+m_{\mathrm{B}^0}^2q^2}{2m_{\mathrm{B}^0}m_{\mathrm{D}^+}},$$ (1) and ranges from 1, when the $`\mathrm{D}^+`$ is produced at rest in the $`\overline{\mathrm{B}}^0`$ rest frame, to about 1.50. Using HQET, the differential partial width for this decay is given by $$\frac{\mathrm{d}\mathrm{\Gamma }(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })}{\mathrm{d}\omega }=\frac{1}{\tau _{\mathrm{B}^0}}\frac{\mathrm{d}\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })}{\mathrm{d}\omega }=𝒦(\omega )^2(\omega )|V_{\mathrm{cb}}|^2,$$ (2) where $`𝒦(\omega )`$ is a known phase space term and $`(\omega )`$ is the hadronic form factor for this decay. Although the shape of this form factor is not known, its magnitude at zero recoil, $`\omega =1`$, can be estimated using HQET. In the heavy quark limit ($`m_\mathrm{b}\mathrm{}`$), $`(\omega )`$ coincides with the Isgur-Wise function which is normalised to unity at the point of zero recoil. Corrections to $`(1)`$ have been calculated to take into account the effects of finite quark masses and QCD corrections, yielding the value and theoretical uncertainty $`(1)=0.913\pm 0.042`$ . Since the phase space factor $`𝒦(w)`$ tends to zero as $`\omega 1`$, the decay rate vanishes at $`\omega =1`$ and the accuracy of the extrapolation relies on achieving a reasonably constant reconstruction efficiency in the region near $`\omega =1`$. Previous measurements of $`|V_{\mathrm{cb}}|`$ have been made using B mesons produced on the $`\mathrm{{\rm Y}}(4S)`$ resonance and in $`\mathrm{Z}^0`$ decays . These analyses used a linear or constrained quadratic expansion of $`(\omega )`$ around $`\omega =1`$. An improved theoretical analysis, based on dispersive bounds and including higher order corrections, has since become available . This results in a parameterisation for $`(\omega )`$ in terms of $`(1)`$ and a single unknown parameter $`\rho ^2`$ constrained to lie in the range $`0.14<\rho ^2<1.54`$, $`\rho ^2`$ corresponding to the slope of $`(\omega )`$ at zero recoil. The previous OPAL measurement used the decay chain $`\mathrm{D}^+\mathrm{D}^0\pi ^+`$, with the $`\mathrm{D}^0`$ meson being reconstructed in the exclusive decay channels $`\mathrm{D}^0\mathrm{K}^{}\pi ^+`$ and $`\mathrm{D}^0\mathrm{K}^{}\pi ^+\pi ^0`$. In this paper, a new analysis is described in which only the $`\pi ^+`$ from the $`\mathrm{D}^+`$ decay is identified, and no attempt is made to reconstruct the $`\mathrm{D}^0`$ decay exclusively. This technique, first employed by DELPHI , gives a much larger sample of $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ decays than the previous measurement, but also larger background, requiring a rather more complex analysis. The measurement of is also updated to use the new parameterisation of $`(\omega )`$, and improved background models and physics inputs. In both cases, the initial number of $`\mathrm{B}^0`$ mesons is determined from other measurements of $`\mathrm{B}^0`$ production in $`\mathrm{Z}^0`$ decays. The new reconstruction technique is described in Section 2, the determination of $`\omega `$ for each event in Section 3, the fit to extract $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ in Section 4 and the systematic errors in Section 5. The updated exclusive measurement is discussed in Section 6 and the measurements are combined and conclusions drawn in Section 7. ## 2 Inclusive reconstruction of $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ events The OPAL detector is well described elsewhere . The data sample used in this analysis consists of about 4 million hadronic $`\mathrm{Z}^0`$ decays collected during the period 1991–1995, at centre-of-mass energies in the vicinity of the $`\mathrm{Z}^0`$ resonance. Corresponding simulated event samples were generated using JETSET 7.4 as described in . Hadronic $`\mathrm{Z}^0`$ decays were selected using standard criteria . To ensure the event was well contained within the acceptance of the detector, the thrust axis direction<sup>2</sup><sup>2</sup>2A right handed coordinate system is used, with positive $`z`$ along the electron beam direction and $`x`$ pointing to the centre of the LEP ring. The polar and azimuthal angles are denoted by $`\theta `$ and $`\varphi `$. was required to satisfy $`|\mathrm{cos}\theta _T|<0.9`$. Charged tracks and electromagnetic calorimeter clusters with no associated tracks were then combined into jets using a cone algorithm , with a cone half angle of 0.65 rad and a minimum jet energy of 5 GeV. The transverse momentum $`p_t`$ of each track was defined relative to the axis of the jet containing it, where the jet axis was calculated including the momentum of the track. A total of 3 117 544 events passed the event selection. The reconstruction of $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ events was then performed by combining high $`p`$ and $`p_t`$ lepton (electron or muon) candidates with oppositely charged pions from the $`\mathrm{D}^+\mathrm{D}^0\pi ^+`$ decay. Electrons were identified and photon conversions rejected using neural network algorithms , and muons were identified as in . Both electrons and muons were required to have momenta $`p>2\mathrm{GeV}`$, transverse momenta with respect to the jet axis $`p_t>0.7\mathrm{GeV}`$, and to lie in the polar angle region $`|\mathrm{cos}\theta |<0.9`$. The event sample was further enhanced in semileptonic b decays by requiring a separated secondary vertex with decay length significance $`L/\sigma _L>2`$ in any jet of the event. The vertex reconstruction algorithm and decay length significance calculation are described fully in . Together with the lepton selection, these requirements result in a sample which is about 90 % pure in $`\mathrm{b}\overline{\mathrm{b}}`$ events. An attempt was made to estimate the $`\mathrm{D}^0`$ direction in each jet containing a lepton candidate. Each track (apart from the lepton) and calorimeter cluster in the jet was assigned a weight corresponding to the estimated probability that it came from the $`\mathrm{D}^0`$ decay. The track weight was calculated from an artificial neural network, trained to separate tracks from b decays and fragmentation tracks in b jets . The network inputs are the track momentum, transverse momentum with respect to the jet axis, and impact parameter significances with respect to the reconstructed primary and secondary vertices (if existing). The cluster weights were calculated using their energies and angles with respect to the jet axis alone, the energies first being corrected by subtracting the energy of any charged tracks associated to the cluster . Beginning with the track or cluster with the largest weight, tracks and clusters were then grouped together until the invariant mass of the group (assigning tracks the pion mass and clusters zero mass) exceeded the charm hadron mass, taken to be 1.8 GeV. If the final invariant mass exceeded 2.3 GeV, the jet was rejected, since Monte Carlo studies showed such high mass $`\mathrm{D}^0`$ candidates were primarily background. For surviving jets, the momentum $`𝐩_{\mathrm{D}^0}`$ of the group was used as an estimate of the $`\mathrm{D}^0`$ direction, giving RMS angular resolutions of about 45 mrad in $`\varphi `$ and $`\theta `$. The $`\mathrm{D}^0`$ energy was calculated as $`E_{\mathrm{D}^0}=\sqrt{p_{\mathrm{D}^0}^2+m_{\mathrm{D}^0}^2}`$. The selection of pions from $`\mathrm{D}^+`$ decays relies on the small mass difference of only 145 MeV between the $`\mathrm{D}^+`$ and $`\mathrm{D}^0`$, which means the pions have very little transverse momentum with respect to the $`\mathrm{D}^0`$ direction. Each track in the jet (other than the lepton) was considered as a slow pion candidate, provided it satisfied $`0.5<p<2.5`$ GeV and had a transverse momentum with respect to the $`\mathrm{D}^0`$ direction of less than 0.3 GeV. If the pion under consideration was included in the reconstructed $`\mathrm{D}^0`$, it was removed and the $`\mathrm{D}^0`$ momentum and energy recalculated. The final selection was made using the reconstructed mass difference $`\mathrm{\Delta }m`$ between $`\mathrm{D}^+`$ and $`\mathrm{D}^0`$ mesons, calculated as $$\mathrm{\Delta }m=\sqrt{E_\mathrm{D}^{}^2|𝐩_\mathrm{D}^{}|^2}m_{\mathrm{D}^0},$$ where the $`\mathrm{D}^+`$ energy is given by $`E_\mathrm{D}^{}=E_{\mathrm{D}^0}+E_\pi `$ and momentum by $`𝐩_\mathrm{D}^{}=𝐩_{\mathrm{D}^0}+𝐩_\pi `$. A new secondary vertex was then iteratively reconstructed around an initial seed vertex formed by the intersection of the lepton and slow pion tracks. Every other track in the jet was added in turn to the seed vertex, and the vertex refitted. The track resulting in the lowest vertex fit $`\chi ^2`$ was retained in the seed vertex for the next iteration. The procedure was repeated until no more tracks could be added without reducing the vertex fit $`\chi ^2`$ probability to less than 1 %. The decay length $`L^{}`$ between the primary vertex and this secondary vertex, and the associated error $`\sigma _L^{}`$, were calculated as in . The pion candidate was accepted if the decay length satisfied $`0.1\mathrm{cm}<L^{}<2`$ cm and the decay length significance satisfied $`L^{}/\sigma _L^{}>2`$. The resulting distributions of $`\mathrm{\Delta }m`$ for opposite and same sign lepton-pion combinations are shown in Figure 1(a) and (b). The predictions of the Monte Carlo simulation are also shown, broken down into contributions from signal $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ events, ‘resonant’ background containing real leptons and slow pions from $`\mathrm{D}^+`$ decay, and combinatorial background, made up of events with fake slow pions, fake leptons or both. In Monte Carlo simulation, about 45 % of opposite sign events with $`\mathrm{\Delta }m<0.17\mathrm{GeV}`$ are signal $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ events, 14 % are resonant background and 41 % are combinatorial background. The resonant background is made up mainly of $`\mathrm{B}^{}\mathrm{D}^+\pi ^{}\mathrm{}^{}\overline{\nu }`$, $`\overline{\mathrm{B}}^0\mathrm{D}^+\pi ^0\mathrm{}^{}\overline{\nu }`$ and $`\overline{\mathrm{B}}_\mathrm{s}\mathrm{D}^+\mathrm{K}^0\mathrm{}^{}\overline{\nu }`$ decays. These are expected to be dominated by b semileptonic decays involving orbitally excited charm mesons (generically referred to as $`\mathrm{D}^{}`$), e.g. $`\mathrm{B}^{}\mathrm{D}^0\mathrm{}^{}\overline{\nu }`$ followed by $`\mathrm{D}^0\mathrm{D}^+\pi ^{}`$. These decays will be denoted collectively by $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$. Small contributions are also expected from $`\mathrm{b}\mathrm{D}^+\tau \overline{\nu }\mathrm{X}`$ decays with the $`\tau `$ decaying leptonically, and $`\mathrm{b}\mathrm{D}^+\mathrm{D}_\mathrm{s}^{}\mathrm{X}`$ with the $`\mathrm{D}_\mathrm{s}^{}`$ decaying semileptonically (each about 1 % of opposite sign events). For same sign events with $`\mathrm{\Delta }m<0.17`$GeV, there is a small resonant contribution of about 6 % from events with a real $`\mathrm{D}^+\mathrm{D}^0\pi ^+`$ where the $`\mathrm{D}^0`$ decays semileptonically, and the rest is combinatorial background. The most important background, from $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ decays, comes from both charged $`\mathrm{B}^+`$ and neutral $`\mathrm{B}^0`$ and $`\mathrm{B}_\mathrm{s}`$ decays, whereas the signal comes only from $`\mathrm{B}^0`$ decays. Therefore the charge $`Q_{\mathrm{vtx}}`$ of the reconstructed secondary vertex containing the lepton and slow pion and its estimated error $`\sigma _{Q_{\mathrm{vtx}}}`$ were calculated, using $`Q_{\mathrm{vtx}}`$ $`=`$ $`{\displaystyle \underset{i}{}}w_iq_i,`$ $`\sigma _{Q_{\mathrm{vtx}}}^2`$ $`=`$ $`{\displaystyle \underset{i}{}}w_i(1w_i)q_i^2,`$ where $`w_i`$ is the weight for track $`i`$ of charge $`q_i`$ to come from the secondary vertex, and the sums are taken over all tracks in the jet . The weights were calculated in a similar way to those used for the $`\mathrm{D}^0`$ direction reconstruction, using a neural network with the track momentum, transverse momentum and impact parameter significances with respect to the reconstructed primary and secondary vertices as inputs. The weights for the lepton and slow pion candidate tracks were set to one. The vertex charge distributions for opposite sign events with $`\mathrm{\Delta }m<0.17`$GeV are shown in Figure 1(c) and (d). The reconstructed vertex charge and error were used to divide the data into different classes enhanced or depleted in $`\mathrm{B}^+`$ decays, thus reducing the effect of this background and increasing the statistical sensitivity. Three classes $`c`$ were used—class 1 where the charge is measured poorly ($`\sigma _{Q_{\mathrm{vtx}}}>0.9`$), class 2 where the charge is measured well and is compatible with a neutral vertex ($`\sigma _{Q_{\mathrm{vtx}}}<0.9`$, $`|Q_{\mathrm{vtx}}|<0.5`$) and class 3 where the charge is measured well and is compatible with a charged vertex ($`\sigma _{Q_{\mathrm{vtx}}}<0.9`$, $`|Q_{\mathrm{vtx}}|>0.5`$). ## 3 Reconstruction of $`𝝎`$ The recoil variable $`\omega `$ was estimated in each event using the reconstructed four-momentum transfer to the $`\mathrm{}\overline{\nu }`$ system: $$q^2=(E_{\mathrm{B}^0}E_\mathrm{D}^{})^2(𝐩_{\mathrm{B}^0}𝐩_\mathrm{D}^{})^2$$ together with equation 1. The $`\mathrm{B}^0`$ and $`\mathrm{D}^+`$ energies $`E_{\mathrm{B}^0}`$ and $`E_\mathrm{D}^{}`$ were estimated directly, whilst the momentum vectors $`𝐩_{\mathrm{B}^0}`$ and $`𝐩_\mathrm{D}^{}`$ were estimated using the energies together with the reconstructed polar and azimuthal angles, as described in more detail below. Since the slow pion has very little momentum in the rest frame of the decaying $`\mathrm{D}^+`$, the momentum (and hence energy) of the $`\mathrm{D}^+`$ was estimated by scaling the reconstructed slow pion momentum by $`m_{\mathrm{D}^+}/m_\pi `$, as for the $`\mathrm{D}^0\mathrm{K}^{}\pi ^+\pi ^0`$ channel in . A small correction (never exceeding 12 %) was applied, as a function of $`\mathrm{cos}\theta ^{}`$, the angle of the slow pion in the rest frame of the $`\mathrm{D}^0`$. This procedure gave a fractional $`\mathrm{D}^+`$ energy resolution of 15 %. The polar and azimuthal angles of the $`\mathrm{D}^+`$ were reconstructed using weighted averages of the slow pion and $`\mathrm{D}^0`$ directions, giving resolutions of about 22 mrad on both $`\varphi `$ and $`\theta `$. The energy of the $`\mathrm{B}^0`$ was estimated using a technique similar to that described in , exploiting the overall energy and momentum conservation in the event to calculate the energy of the unreconstructed neutrino. First, the energy $`E_{\mathrm{bjet}}`$ of the jet containing the $`\mathrm{B}^0`$ was inferred from the measured particles in the rest of the event, by treating the event as a two-body decay of a $`\mathrm{Z}^0`$ into a b jet (whose mass was approximated by the $`\mathrm{B}^0`$ mass) and another object making up the rest of the event. Then, the total fragmentation energy $`E_{\mathrm{frag}}`$ in the b jet was estimated from the measured visible energy in the b jet and the identified $`\mathrm{B}^0`$ decay products: $`E_{\mathrm{frag}}=E_{\mathrm{vis}}E_{\mathrm{}}E_\mathrm{D}^{}`$. Finally, the $`\mathrm{B}^0`$ energy was calculated as $`E_{\mathrm{B}^0}=E_{\mathrm{bjet}}E_{\mathrm{frag}}`$, giving an RMS resolution of 4.4 GeV. The b direction was estimated using a combination of two techniques. In the first, the momentum vector of the $`\mathrm{B}^0`$ was reconstructed from its decay products: $$𝐩_{\mathrm{B}^0}=𝐩_\mathrm{D}^{}+𝐩_{\mathrm{}}+𝐩_{\overline{\nu }},$$ the neutrino energy being estimated from the reverse of the event visible momentum vector: $`𝐩_{\overline{\nu }}=𝐩_{\mathrm{vis}}`$. The visible momentum was calculated using all the reconstructed tracks and clusters in the event, with a correction for charged particles measured both in the tracking detectors and calorimeters . This direction estimate is strongly degraded if a second neutrino is present (e.g. from another semileptonic b decay in the opposite hemisphere of the event), and its error was parameterised as a function of the visible energy in the opposite hemisphere. The resulting resolution is typically between 40 and 100 mrad for both $`\varphi `$ and $`\theta `$. The second method of estimating the $`\mathrm{B}^0`$ direction used the vertex flight direction—i.e. the direction of the vector between the primary vertex and the secondary vertex reconstructed around the lepton and slow pion as described in Section 2. The accuracy of this estimate is strongly dependent on the $`\mathrm{B}^0`$ decay length, and was used only if the decay length significance $`L^{}/\sigma _L^{}`$ exceeded 3. After this cut, the angular resolution varies between about 15 and 100 mrad, and is worse for $`\theta `$ in the 1991 and 1992 data where accurate $`z`$ information from the silicon microvertex detector was not available. The $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ candidate was rejected if the two reconstruction methods gave $`\theta `$ or $`\varphi `$ angles disagreeing by more than three standard deviations, which happened in 7 % of Monte Carlo signal events. Finally, the two $`\mathrm{B}^0`$ direction estimates were combined according to their estimated uncertainties, giving average resolutions of 35 mrad on $`\varphi `$ and 43 mrad on $`\theta `$, including events where only the first method was used. The estimate of $`q^2`$ derived from the $`\mathrm{B}^0`$ and $`\mathrm{D}^+`$ energies and angles was improved by applying the constraint that the mass $`\mu `$ of the neutrino produced in the $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ decay should be zero. The neutrino mass is given from the reconstructed quantities by: $$\mu ^2=(E_{\mathrm{B}^0}E_\mathrm{D}^{}E_{\mathrm{}})^2(𝐩_{\mathrm{B}^0}𝐩_\mathrm{D}^{}𝐩_{\mathrm{}})^2.$$ The constraint was implemented by calculating $`q^2`$ using a kinematic fit, incorporating the measured values and estimated uncertainties of the $`\mathrm{B}^0`$ and $`\mathrm{D}^+`$ energies and angles (the $`\mathrm{B}^0`$ angular uncertainties varying event by event). This procedure improved the average $`q^2`$ resolution from 2.78 GeV<sup>2</sup> to 2.57 GeV<sup>2</sup>. In Monte Carlo simulation, 11 % of signal events were reconstructed with $`q^2<0`$, corresponding to an unphysical value of $`\omega `$ larger than $`\omega _{\mathrm{max}}1.5`$, and were rejected. The resulting distributions of reconstructed $`\omega `$ for various ranges of true $`\omega `$ (denoted $`\omega ^{}`$) in simulated $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ decays are shown in Figure 2(a–e). The average RMS $`\omega `$ resolution is about 0.12, but there are significant non-Gaussian tails. The resolution was parameterised (separately for the 1991–2 and 1993–5 data) as a continuous function $`R(\omega ,\omega ^{})`$ giving the expected distribution of reconstructed $`\omega `$ for each true value $`\omega ^{}`$. The resolution function was implemented as the sum of two asymmetric Gaussians (i.e. with different widths either side of the peak) whose parameters were allowed to vary as a function of $`\omega ^{}`$. The convolution of this resolution function with the Monte Carlo $`\omega ^{}`$ distribution is also shown in Figure 2(a–e), demonstrating that the resolution function models the $`\omega `$ distributions well. The efficiency to reconstruct $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ decays $`ϵ(\omega ^{})`$ is shown in Figure 2(f), together with a second order polynomial parameterisation. The efficiency varies with $`\omega ^{}`$, but is reasonably flat in the critical region near $`\omega ^{}=1`$ where the extrapolation to measure $`(1)|V_{\mathrm{cb}}|`$ is carried out. ## 4 Fit and results The values of $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ were extracted using an extended maximum likelihood fit to the reconstructed mass difference $`\mathrm{\Delta }m`$, recoil $`\omega `$ and vertex charge class $`c`$ of each event. Both opposite and same sign events with $`\mathrm{\Delta }m<0.3\mathrm{GeV}`$ and $`\omega <\omega _{\mathrm{max}}`$ were used in the fit, the high $`\mathrm{\Delta }m`$ and same sign events serving to constrain the combinatorial background normalisation and shapes in the opposite sign low $`\mathrm{\Delta }m`$ region populated by the $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ decays. Using the $`\mathrm{\Delta }m`$ value from each event in the fit, rather than just dividing the data into low $`\mathrm{\Delta }m`$ ‘signal’ and high $`\mathrm{\Delta }m`$ ‘sideband’ mass regions, increases the statistical sensitivity as the signal purity varies considerably within the low $`\mathrm{\Delta }m`$ region. The logarithm of the overall likelihood was given by $$\mathrm{ln}=\underset{i=1}{\overset{M^a}{}}\mathrm{ln}_i^a+\underset{j=1}{\overset{M^b}{}}\mathrm{ln}_j^bN^aN^b$$ (3) where the sums of individual event log-likelihoods $`\mathrm{ln}_i^a`$ and $`\mathrm{ln}_j^b`$ are taken over all the observed $`M^a`$ opposite sign and $`M^b`$ same sign events in the data sample, and $`N^a`$ and $`N^b`$ are the corresponding expected numbers of events. The likelihood for each opposite sign event was given in terms of different types or sources of event by $$_i^a(\mathrm{\Delta }m_i,\omega _i,c_i)=\underset{s=1}{\overset{4}{}}N_s^af_{s,c_i}^aM_{s,c_i}(\mathrm{\Delta }m_i)P_s(\omega _i)$$ (4) where $`N_s^a`$ is the number of expected events for source $`s`$, $`f_{s,c}^a`$ is the fraction of events in source $`s`$ appearing in vertex charge class $`c`$, $`M_{s,c}(\mathrm{\Delta }m)`$ is the mass difference distribution for source $`s`$ in class $`c`$ and $`P_s(\omega )`$ the recoil distribution for source $`s`$. For each source, the vertex charge fractions $`f_{s,c}^a`$ sum to one and the mass difference $`M_{s,c}(\mathrm{\Delta }m_i)`$ and recoil $`P_s(\omega )`$ distributions are normalised to one. The total number of expected events is given by the sum of the individual contributions: $`N^a=_{s=1}^4N_s^a`$. There are four opposite sign sources: (1) signal $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ events, (2) $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ events where the $`\mathrm{D}^+`$ is produced via an intermediate resonance ($`\mathrm{D}^{}`$), (3) other opposite sign background involving a genuine lepton and a slow pion from $`\mathrm{D}^+`$ decay and (4) combinatorial background. The sum of sources 2 and 3 are shown as ‘resonant background’ in Figure 1. A similar expression to equation 4 was used for $`_j^b`$, the event likelihood for same sign events. In this case, only sources 3 and 4 contribute. The mass difference distributions $`M_{c,s}(\mathrm{\Delta }m)`$ for sources 1–3 were represented by analytic functions, whose parameters were determined using large numbers of simulated events, as were the recoil distributions $`P_s(\omega )`$ for sources 2 and 3. The fractions in each vertex charge class for sources 1–3 were also taken from simulation. For the signal (source 1), the product of the expected number of events $`N_1^a`$ and recoil distribution $`P_1(\omega )`$ was given by convolving the differential partial decay width (equation 2) with the signal resolution function and reconstruction efficiency: $$N_1^aP_1(\omega )=4N_\mathrm{Z}R_\mathrm{b}f_{\mathrm{B}^0}\tau _{\mathrm{B}^0}\mathrm{Br}(\mathrm{D}^+\mathrm{D}^0\pi ^+)_1^{\omega _{\mathrm{max}}}^2(\omega ^{})|V_{\mathrm{cb}}|^2𝒦(\omega ^{})ϵ(\omega ^{})R(\omega ,\omega ^{})𝑑\omega ^{}$$ where $`N_\mathrm{Z}`$ is the number of hadronic $`\mathrm{Z}^0`$ decays passing the event selection, $`R_\mathrm{b}\mathrm{\Gamma }_{\mathrm{b}\overline{\mathrm{b}}}/\mathrm{\Gamma }_{\mathrm{had}}`$ is the fraction of hadronic $`\mathrm{Z}^0`$ decays to $`\mathrm{b}\overline{\mathrm{b}}`$, $`f_{\mathrm{B}^0}`$ the fraction of b quarks hadronising to a $`\overline{\mathrm{B}}^0`$ and $`\tau _{\mathrm{B}^0}`$ the $`\mathrm{B}^0`$ lifetime. The factor of four accounts for the two b hadrons produced per $`\mathrm{Z}^0\mathrm{b}\overline{\mathrm{b}}`$ event and the two identified lepton species (electrons and muons). The form factor $`(\omega ^{})`$ is given in in terms of the normalisation $`(1)`$ and slope parameter $`\rho ^2`$. The efficiency function $`ϵ(\omega ^{})`$ and resolution function $`R(\omega ,\omega ^{})`$ were described in Section 3, and the known phase space factor $`𝒦(\omega ^{})`$ is given in . The assumed values of the numerical quantities are given in Table 1. Since the data are divided into different vertex charge classes enhanced and depleted in $`\mathrm{B}^+`$ decays, the fit gives some information on the amount of $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ background. The predicted level of this background in the fit was therefore allowed to vary under a Gaussian constraint corresponding to the branching ratio of $`\mathrm{Br}(b\mathrm{D}^+\mathrm{h}\mathrm{}\overline{\nu })=(0.76\pm 0.16)\%`$. The latter has been calculated from the measured branching ratio $`\mathrm{Br}(\mathrm{b}\mathrm{D}^+\pi ^{}\mathrm{}\overline{\nu }\mathrm{X})=(0.473\pm 0.095)\%`$ , assuming isospin and SU(3) flavour symmetry to obtain the corresponding $`\mathrm{b}\mathrm{D}^+\pi ^0\mathrm{}\overline{\nu }`$ and $`\mathrm{b}\mathrm{D}^+\mathrm{K}^0\mathrm{}\overline{\nu }`$ branching ratios. A scaling factor of $`0.75\pm 0.25`$ was included for the last branching ratio to account for possible SU(3) violation effects reducing the branching ratio $`\mathrm{D}_\mathrm{s}^+\mathrm{D}^+\mathrm{K}^0`$ compared to the expectation of $`\frac{3}{2}\mathrm{Br}(\mathrm{D}^+\mathrm{D}^+\pi ^0)`$. The $`P_2(\omega )`$ distribution for these events was taken from simulation, using the calculation of Leibovich et al. to predict their recoil spectrum. The numbers $`N_3^{a,b}`$ and $`P_3(\omega )`$ distributions for the small background contributions covered by source 3 (both opposite and same sign) were taken from Monte Carlo simulation, with branching ratios adjusted to the values given in Table 1. The branching ratio for $`\mathrm{b}\mathrm{D}^+\tau ^{}\overline{\nu }`$ was derived using the inclusive branching ratio $`\mathrm{Br}(\mathrm{b}\tau \mathrm{X})=(2.6\pm 0.4)\%`$ , assuming a $`\mathrm{D}^+`$ is produced in a fraction $`0.25\pm 0.03`$ of the time, as seen for the corresponding decays $`\mathrm{b}\mathrm{D}^+\mathrm{}\mathrm{X}`$ and $`\mathrm{b}\mathrm{}\mathrm{X}`$ ($`\mathrm{}=\mathrm{e},\mu `$) . The rate of $`\mathrm{b}\mathrm{D}^+\mathrm{D}_\mathrm{s}^{}\mathrm{X}`$ was assumed to be dominated by the two body decay $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{D}_\mathrm{s}^{()}`$. The rate of real $`\mathrm{D}^+`$ decays in the same sign background depends on the production fractions of $`\mathrm{D}^+`$ in $`\mathrm{b}\overline{\mathrm{b}}`$ and $`\mathrm{c}\overline{\mathrm{c}}`$ events, which were taken from . The parameters of the analytic functions describing the combinatorial background ($`N_4^a`$, $`N_4^b`$, $`f_{4,c}^a`$, $`f_{4,c}^b`$, $`M_{4,c}(\mathrm{\Delta }m)`$ and $`P_4(\omega )`$) were fitted entirely from the data, with only the choice of functional forms motivated by simulation. The shapes of the mass and recoil functions (including a small correlation between $`\mathrm{\Delta }m`$ and $`\omega `$) are constrained by the same sign sample (which is almost entirely combinatorial background), and are the same for each vertex charge class $`c`$. The opposite sign high $`\mathrm{\Delta }m`$ region serves to normalise the number of combinatorial background events in the low $`\mathrm{\Delta }m`$ region for each vertex charge class. The values of $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ were extracted by maximising the total likelihood given by equation 3. The values of $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ were allowed to vary, together with the level of $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ background and 13 auxiliary parameters describing the combinatorial background distributions. A result of $`(1)|V_{\mathrm{cb}}|`$ $`=`$ $`(37.5\pm 1.2)\times 10^3,`$ $`\rho ^2`$ $`=`$ $`1.12\pm 0.14`$ was obtained, where the errors are only statistical. The correlation between $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ is 0.77. The distributions of reconstructed $`\omega `$ for opposite and same sign events with $`\mathrm{\Delta }m<0.17`$ GeV, together with the fit results, are shown in Figure 3. The fit describes the data well, both in this region and the high $`\mathrm{\Delta }m`$ region dominated by combinatorial background, for all three of the vertex charge classes. By integrating the differential partial decay width (equation 2) over all values of $`\omega `$, the branching ratio $`\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })`$ was also determined to be $$\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })=(5.92\pm 0.27)\%,$$ where again the error is statistical only. This result is consistent with the world average once systematic errors are included. Many previous results have been obtained using a constrained quadratic expansion for the form factor: $`(\omega )=(1)\left[1a^2(\omega 1)+b(\omega 1)^2\right]`$, where $`a`$ is a slope parameter to be determined by the fit, and $`b`$ is constrained to $`b=0.66a^20.11`$ . To allow comparison with such measurements, the fit was also performed with this parameterisation of $`(\omega )`$, giving the results $`(1)|V_{\mathrm{cb}}|=(36.9\pm 1.2)\times 10^3`$ and $`a^2=0.88\pm 0.14`$, the correlation between $`(1)|V_{\mathrm{cb}}|`$ and $`a^2`$ being 0.79. The difference in the two curvature parameters $`\rho ^2`$ and $`a^2`$ is in good agreement with the expectation of $`\rho ^2a^20.21`$ . ## 5 Systematic Errors Systematic errors arise from the uncertainties in the fit input parameters given in Table 1, the Monte Carlo modelling of the signal $`\omega `$ resolution, the recoil spectrum of $`\mathrm{b}\mathrm{D}^{}\mathrm{}\overline{\nu }`$ decays and selection efficiencies, and possible biases in the fitting method. The resulting systematic errors on the values of $`(1)|V_{\mathrm{cb}}|`$, $`\rho ^2`$ and $`\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })`$ are summarized in Table 2 and described in more detail below. The various numerical fit inputs were each varied according to the errors given in Table 1 and the fit repeated to assess the resulting uncertainties. The calculation of Leibovich et al. was used to simulate the recoil spectrum of $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ decays, assumed to be produced via the semileptonic decay $`\mathrm{B}^0\mathrm{D}^{}\mathrm{}\overline{\nu }`$. Here $`\mathrm{D}^{}`$ represents a P-wave orbitally excited charm meson. The calculation predicts the recoil spectra and relative rates of semileptonic decays involving both the narrow $`\mathrm{D}_1`$ and $`\mathrm{D}_2^{}`$ states and wide $`\mathrm{D}_0^{}`$ and $`\mathrm{D}_1^{}`$ states. All these decays are suppressed close to $`\omega =1`$ by an extra factor of $`(\omega ^21)`$ when compared with the signal $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ decays. This reduces the uncertainty due to the rate of $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ decays in the extrapolation of the signal recoil spectrum to $`\omega =1`$. Non-resonant $`\overline{\mathrm{B}}\mathrm{D}^+\mathrm{h}\mathrm{}^{}\overline{\nu }`$ decays are not included in the model, but are not expected to contribute close to $`\omega =1`$. The differential decay rates in are given in terms of five possible expansion schemes and several unknown parameters: a kinetic energy term $`\eta _{\mathrm{ke}}`$ and the slopes of the Isgur-Wise functions for the narrow and wide $`\mathrm{D}^{}`$ states $`\widehat{\tau _1}`$ and $`\widehat{\zeta _1}`$. These parameters were varied within the allowed ranges $`0.75<\eta _{\mathrm{ke}}<0.75`$GeV, $`2<\widehat{\tau _1}<1`$ and $`2<\widehat{\zeta _1}<0`$ subject to the constraint that the ratio $`R=\mathrm{\Gamma }(\overline{B}D_2^{}\mathrm{}\overline{\nu })/(\mathrm{\Gamma }(\overline{B}D_1\mathrm{}\overline{\nu })`$ lie within the measured range $`R=0.37\pm 0.16`$ . This excludes the expansion schemes $`A_{\mathrm{}}`$ and $`B_{\mathrm{}}`$ of and constrains the allowable values of $`\eta _{\mathrm{ke}}`$ in the others. The fraction of $`\overline{\mathrm{B}}\mathrm{D}^{}\mathrm{}^{}\overline{\nu }`$ decays involving the narrow $`\mathrm{D}_1`$ and $`\mathrm{D}_2^{}`$ states, which is not precisely predicted, was varied in the range $`0.22\pm 0.06`$, obtained by comparing the measured rates for $`\mathrm{B}^0`$ semileptonic decays involving $`\mathrm{D}^+`$, $`\mathrm{D}^+`$, $`\mathrm{D}_1`$ and $`\mathrm{D}_2^{}`$ with the inclusive semileptonic decay rate . The systematic errors were determined as half the difference between the two parameter sets giving the most extreme variations in $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$, and the central values were adjusted to half way between these two extremes. The values of both $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ are most sensitive to variations in $`\eta _{\mathrm{ke}}`$, which is constrained by the measured value of $`R`$. The effect of uncertainties in the average b hadron energy $`x_E=E_\mathrm{b}/E_{\mathrm{beam}}`$ was assessed in Monte Carlo simulation by reweighting the events so as to vary $`x_E`$ in the range $`0.702\pm 0.008`$ , and repeating the fit. The largest of the variations observed using the fragmentation functions of Peterson, Collins and Spiller, Kartvelishvili and the Lund group were taken as systematic errors. The signal reconstruction efficiency and vertex charge distributions are sensitive to the $`\mathrm{B}^0`$ decay multiplicity, which depends only on the $`\mathrm{D}^0`$ decay for the $`\mathrm{B}^0`$ decay channels of interest. The systematic error was assessed by varying separately the $`\mathrm{D}^0`$ charged and $`\pi ^0`$ decay multiplicities in Monte Carlo simulation according to the measurements of Mark III . The branching ratio $`\mathrm{D}^0\mathrm{K}^0,\overline{\mathrm{K}}^0`$ was also varied according to its uncertainty . The resulting uncertainties on $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ from each variation were added in quadrature to determine the total systematic errors. The modelling of the $`\omega `$ resolution depends on the description of the $`\mathrm{D}^+`$ and $`\mathrm{B}^0`$ energy and angular distributions in the simulation. The reconstructed $`\mathrm{D}^+`$ and $`\mathrm{B}^0`$ energy distributions in data and simulation were compared, and the means were found to differ by 0.04 and 0.13 GeV respectively. The opposite hemisphere missing energy was found to agree within 5 %. The corresponding systematics were assessed by shifting or scaling the data distributions and repeating the $`\omega `$ reconstruction and fit. The modelling of the angular resolution was checked by studying the agreement of the two angular estimators—the slow pion and $`\mathrm{D}^0`$ directions for the the $`\mathrm{D}^+`$, and the missing energy vector and vertex flight directions for the $`\mathrm{B}^0`$. The angular resolutions were found to be up to 5 % worse in data, and the systematic error was assessed by degrading the simulated resolution appropriately. Finally, the fraction of events with $`\omega `$ reconstructed in the physical region ($`\omega <\omega _{\mathrm{max}}`$) was found to be 3.5 % smaller in the data, in both opposite and same sign charge samples. The reconstruction efficiency was corrected for this effect, and an additional systematic error of half the correction ($`1.7\%`$) assumed. The final systematic errors due to $`\omega `$ resolution modelling are dominated by the $`\mathrm{B}^0`$ $`\theta `$ resolution. The electron identification efficiency has been studied using control samples of pure electrons from $`\mathrm{e}^+\mathrm{e}^{}\mathrm{e}^+\mathrm{e}^{}`$ events and photon conversions, and found to be modelled to a a precision of 4.1 % . The muon identification efficiency has been studied using muon pairs producted in two-photon collisions and $`\mathrm{Z}^0\mu ^+\mu ^{}`$ events, giving an uncertainty of 2.1 % . The fraction of hemispheres with identified leptons which also had a selected secondary vertex was found to be about 4 % less in data than in simulation. The overall fraction of vertex tagged hemispheres was also found to be about 4 % lower in data. These discrepancies were translated into systematic errors on the efficiency to tag a semileptonic b decay with a secondary vertex in either the same or the opposite hemisphere, in each case attributing the whole discrepancy to a mismodelling of b hadron decays. The resulting errors on the same and opposite hemisphere tagging efficiencies were taken to be fully correlated. The overall track reconstruction efficiency is known to be modelled to a precision of 1 % , and a similar uncertainty was found to be appropriate for the particular class of slow pion tracks from $`\mathrm{D}^+`$ decays. The systematic error was assessed by randomly removing 1 % of tracks in the simulation and repeating the fit. Uncertainties in the tracking detector resolution affect the efficiency, $`\omega `$ reconstruction and vertex charge distributions. The associated error was assessed in the simulation by applying a global 10 % degradation to all tracks, independently in the $`r`$-$`\varphi `$ and $`r`$-$`z`$ planes, as in . The entire fitting procedure was tested on a fully simulated Monte Carlo sample seven times bigger than the data, with true values of $`(1)|V_{\mathrm{cb}}|=32.5\times 10^3`$ and $`\rho ^2=1.3`$. The fit gave the results $`(1)|V_{\mathrm{cb}}|=(31.8\pm 0.5)\times 10^3`$ and $`\rho ^2=1.25\pm 0.07`$, consistent with the true values. For each variable, the larger of the deviation of the result from the true value and the statistical error were taken as a systematic errors due to possible biases in the fit. Additionally, the large Monte Carlo sample was reweighted to change the values of $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$, and the fit correctly recovered the modified values. To verify the correctness of the statistical errors returned by the fit, it was performed on many separate subsamples, and the distribution of fitted results studied. Further checks on the data included performing the analysis separately for $`\mathrm{B}^0`$ decays involving electrons and muons, dividing the sample according to the year of data taking, and varying the lepton transverse momentum cut. In all cases, consistent results were obtained. Including all systematic uncertaintities, the final result of the inclusive analysis is $`(1)|V_{\mathrm{cb}}|`$ $`=`$ $`(37.5\pm 1.2\pm 2.5)\times 10^3,`$ $`\rho ^2`$ $`=`$ $`1.12\pm 0.14\pm 0.29,`$ where the first error is statistical and the second systematic in each case. ## 6 Measurement using exclusively reconstructed $`𝐃^\mathbf{}\mathbf{+}`$ decays In this analysis, the $`\mathrm{D}^0`$ from the $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$, $`\mathrm{D}^+\mathrm{D}^0\pi ^+`$ decay is reconstructed exclusively in the decay modes $`\mathrm{D}^0\mathrm{K}^{}\pi ^+`$ (‘3-prong’) and $`\mathrm{D}^0\mathrm{K}^{}\pi ^+\pi ^0`$ (‘satellite’—where the $`\pi ^0`$ is not reconstructed). The event selection, reconstruction and determination of $`\omega `$ are exactly the same as described in . The determination of the signal and background fractions, and the fit to extract $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ are performed as in , but have been updated using the improved form factor parameterisation , the updated input parameters given in Table 1 and the $`\mathrm{b}\mathrm{D}^{}\mathrm{}\overline{\nu }`$ decay model discussed in Section 5. The fit is performed on 814 3-prong and 1396 satellite candidates, of which $`505\pm 44`$ and $`754\pm 72`$ are attributed to $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ signal decays. The result of the fit is $`(1)|V_{\mathrm{cb}}|`$ $`=`$ $`(36.8\pm 1.6\pm 2.0)\times 10^3,`$ $`\rho ^2`$ $`=`$ $`1.31\pm 0.21\pm 0.16,`$ where again the first errors are statistical and the second systematic. The statistical correlation between $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ is 0.95. The distribution of reconstructed $`\omega `$ for selected candidates (both 3-prong and satellite) is shown in Figure 4. The branching ratio $`\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })`$ has also been determined to be $$\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })=(5.11\pm 0.19\pm 0.49)\%.$$ The systematic errors arise from uncertainties in the background levels in the selected samples, as well as uncertainties in the Monte Carlo simulations. They have been evaluated following the procedures described in , and are summarised in Table 3. The selection efficiency error includes contributions from lepton identification efficiency, b fragmentation and detector resolution uncertanties, described in detail in . The largest change with respect to the previous result is due to the improved $`\mathrm{b}\mathrm{D}^{}\mathrm{}\overline{\nu }`$ modelling, with the suppression of this background at values of $`\omega `$ close to one. This reduces the statistical error, the systematic uncertainty due to the rate of such decays, and shifts the central value of $`(1)|V_{\mathrm{cb}}|`$ upwards as compared to . ## 7 Conclusions The CKM matrix element $`|V_{\mathrm{cb}}|`$ has been measured by studying the rate of the semileptonic decay $`\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu }`$ as a function of the recoil kinematics of both inclusively and exclusively reconstructed $`\mathrm{D}^+`$ mesons. The two results are combined, taking into account the statistical correlation of 18 % and correlated systematic errors from physics inputs and detector resolution. The results are: $`(1)|V_{\mathrm{cb}}|`$ $`=`$ $`(37.1\pm 1.0\pm 2.0)\times 10^3,`$ $`\rho ^2`$ $`=`$ $`1.21\pm 0.12\pm 0.20,`$ $`\mathrm{Br}(\overline{\mathrm{B}}^0\mathrm{D}^+\mathrm{}^{}\overline{\nu })`$ $`=`$ $`(5.26\pm 0.20\pm 0.46)\%,`$ where the first result is statistical and the second systematic in each case. The statistical and systematic correlations between $`(1)|V_{\mathrm{cb}}|`$ and $`\rho ^2`$ are 0.90 and 0.54 respectively. These results supersede our previous publication . They are consistent with other determinations of $`(1)|V_{\mathrm{cb}}|`$ at LEP and the $`\mathrm{{\rm Y}}(4S)`$ resonance . The branching ratio is consistent with the world average result of $`(4.60\pm 0.27)`$ % . The result for $`(1)|V_{\mathrm{cb}}|`$ is the most precise to date from any single experiment. Using the theoretical estimate $`(1)=0.913\pm 0.042`$ , the value of $`|V_{\mathrm{cb}}|`$ is determined to be $$|V_{\mathrm{cb}}|=(40.7\pm 1.1\pm 2.2\pm 1.6)\times 10^3,$$ where the uncertainties are statistical, systematic and theoretical respectively. ## Acknowledgements We particularly wish to thank the SL Division for the efficient operation of the LEP accelerator at all energies and for their continuing close cooperation with our experimental group. We thank our colleagues from CEA, DAPNIA/SPP, CE-Saclay for their efforts over the years on the time-of-flight and trigger systems which we continue to use. In addition to the support staff at our own institutions we are pleased to acknowledge the Department of Energy, USA, National Science Foundation, USA, Particle Physics and Astronomy Research Council, UK, Natural Sciences and Engineering Research Council, Canada, Israel Science Foundation, administered by the Israel Academy of Science and Humanities, Minerva Gesellschaft, Benoziyo Center for High Energy Physics, Japanese Ministry of Education, Science and Culture (the Monbusho) and a grant under the Monbusho International Science Research Program, Japanese Society for the Promotion of Science (JSPS), German Israeli Bi-national Science Foundation (GIF), Bundesministerium für Bildung, Wissenschaft, Forschung und Technologie, Germany, National Research Council of Canada, Research Corporation, USA, Hungarian Foundation for Scientific Research, OTKA T-029328, T023793 and OTKA F-023259.
warning/0003/math-ph0003032.html
ar5iv
text
# Universal Similarity Factorization Equalities over Real Clifford Algebras Abstract. A variety of universal similarity factorization equalities over real Clifford algebras $`_{p,q}`$ are established. On the basis of these equalities, real, complex and quaternion matrix representations of elements in $`_{p,q}`$ can be explicitly determined. Key words: Clifford algebras; similarity factorization equalities; matrix representations. AMS subject classifications: 15A66, 15A23. 1. INTRODUCTION The aim of this paper is to establish universal similarity factorization equalities between elements of real Clifford algebras $`_{p,q}`$ and matrices with elements in $`,𝒞`$ and $`,`$ where $`,𝒞`$ and $``$ stand for real number field, complex number field and quaternion skew field, respectively. A direct motivation for considering this problem comes from the following basic universal similarity factorization equality for complex numbers: $$\left[\begin{array}{cc}1& \hfill i\\ i& \hfill 1\end{array}\right]\left[\begin{array}{cc}a+bi& 0\\ 0& abi\end{array}\right]\left[\begin{array}{cc}1& \hfill i\\ i& \hfill 1\end{array}\right]=\left[\begin{array}{cc}a& \hfill b\\ b& \hfill a\end{array}\right].$$ $`(1.1)`$ This equality clearly reveals three fundamental facts on the field of complex numbers (a) $`𝒞`$ is algebraically isomorphic to the matrix algebra $`𝒜=\left\{\left[\begin{array}{cc}a& \hfill b\\ b& \hfill a\end{array}\right]\right|a,b\}`$ through the bijective map $`\varphi :a+bi\left[\begin{array}{cc}a& \hfill b\\ b& \hfill a\end{array}\right]`$. (b) Every complex number $`p=a+bi`$ has a faithful matrix representation $`\varphi (p)=\left[\begin{array}{cc}a& \hfill b\\ b& \hfill a\end{array}\right]`$ over the real number field $``$. (c) All real matrices of the form $`\left[\begin{array}{cc}a& \hfill b\\ b& \hfill a\end{array}\right]`$ can uniformly be diagonalized over the complex number field $`𝒞`$. Following Eq.(1.1), a natural equation can directly be asked: Can we extend Eq.(1.1) to any real Clifford algebra $`_{p,q}`$? The answer to this question is positive. In this paper, we shall present a set of general methods for establishing such kinds of universal similarity factorization equalities over all real Clifford algebras $`_{p,q}`$. As is well known, the real Clifford algebra $`_{p,q}`$ is an associative algebra, with identity 1, defined on $`p+q=n`$ generators $`e_1,e_2,\mathrm{},e_n`$ subject to the multiplication laws $$e_i^2=\{\begin{array}{c}+1fori=1,2,\mathrm{},p,\hfill \\ 1fori=p+1,p+2,\mathrm{},p+q=n,\hfill \end{array}$$ $`(1.2)`$ $$e_ie_j+e_je_i=0forij,i,j=1,2,\mathrm{},n.$$ $`(1.3)`$ and $`e_1e_2\mathrm{}e_n\pm 1`$. In that case $`_{p,q}`$ is spanned as a $`2^n`$-dimensional vector space with a basis $`\{e_A\}`$, where the multiindex $`A`$ ranges all naturally ordered subsets of the first positive integer set $`\{1,2,\mathrm{},n\}`$; the basis element $`e_A`$, where $`A=(i_1,i_2,\mathrm{},i_k)`$ with $`1i_1<i_2<\mathrm{}<i_kn`$, is defined as the product $$e_A=e_{(i_1,i_2,\mathrm{},i_k)}e_{i_1}e_{i_2}\mathrm{}e_{i_k}.$$ In particular, $`e_A=1,`$ when $`A=\mathrm{}`$. For simplicity, a brief notation $`e_{[n]}=e_1e_2\mathrm{}e_n`$ will be adopted in the sequel. The square of $`e_{[n]}`$ is $$e_{[n]}^2=(1)^{\frac{1}{2}n(n1)}e_1^2e_2^2\mathrm{}e_n^2.$$ Any element $`a_{p,q}`$ can be expressed as $$a=\underset{A}{}a_Ae_A,a_A,$$ where $`A`$ ranges all naturally ordered subsets of $`\{1,2,\mathrm{},n\}`$. We shall also adopt the following notation in the sequel $$_{p,q}=\{e_1,\mathrm{},e_p,\epsilon _1,\mathrm{},\epsilon _q|e_i^2=1,\epsilon _j^2=1,i=1,\mathrm{},p,j=1,\mathrm{},q\},$$ or simply $$_{p,q}=\{e_1,\mathrm{},e_p,\epsilon _1,\mathrm{},\epsilon _q\},$$ and use $$_{p,q}^{m\times n}=^{m\times n}\{e_1,\mathrm{},e_p,\epsilon _1,\mathrm{},\epsilon _q\}$$ to stand for the collection of all $`m\times n`$ matrices over $`_{p,q}.`$ As to the algebraic structure of the Clifford algebra $`_{p,q}`$ with $`p+q=n`$, it is well-known(see, e.g., , , , ) that $`_{p,q}`$ satisfies the following algebraic isomorphisms $$_{p,q}\{\begin{array}{c}(2^{n/2})ifqp0,6(mod8),\hfill \\ 𝒞(2^{(n1)/2})ifqp1,5(mod8),\hfill \\ (2^{(n2)/2})ifqp2,4(mod8),\hfill \\ {}_{}{}^{2}(2^{(n3)/2})ifqp3(mod8),\hfill \\ {}_{}{}^{2}(2^{(n3)/2})ifqp7(mod8),\hfill \end{array}$$ $`(1.4)`$ where $`(s),𝒞(s),(s)`$ stand for the matrix algebras $`^{s\times s},𝒞^{s\times s},^{s\times s}`$, respectively, and $`{}_{}{}^{2}(s),^2(s)`$ stand for the matrix algebras $${}_{}{}^{2}(s)=\left\{\left[\begin{array}{cc}A& O\\ O& B\end{array}\right]\right|A,B^{s\times s}\},^2(s)=\left\{\left[\begin{array}{cc}A& O\\ O& B\end{array}\right]\right|A,B^{s\times s}\}.$$ For some low values of $`p`$ and $`q`$, Eq.(1.4) can be expressed as $$_{1,0}^2,_{0,1}=𝒞,$$ $`(1.5)`$ $$_{2,0}^{2\times 2},_{1,1}^{2\times 2},_{0,2}=,$$ $`(1.6)`$ $$_{3,0}𝒞^{2\times 2},_{2,1}^2^{2\times 2},_{1,2}𝒞^{2\times 2},_{0,3}^2,$$ $`(1.7)`$ $$_{4,0}^{2\times 2},_{3,1}^{4\times 4},_{2,2}𝒞^{4\times 4},_{1,3}^{2\times 2},_{0,4}^{2\times 2}.$$ $`(1.8)`$ In addition, according to the periodicity theorem on Clifford algebras( see, e.g., and ), it is also well-known that $$_{p+8,q}_{p,q}^{16\times 16},_{p,q+8}_{p,q}^{16\times 16}$$ $`(1.9)`$ hold for all finite $`p`$ and $`q`$. The isomorphisms listed above imply that there exists a one-to-one correspondence, preserving algebraic operations, between elements of $`_{p,q}`$ and matrices with elements in $``$ or $`𝒞`$ or $``$, what we shall do in the present paper is to explicitly establish universal similarity factorization equalities between elements of $`_{p,q}`$ and matrices with elements in $`,`$ $`𝒞`$, and $`.`$ A key tool used in the sequel is given below. Lemma 1.1. Let $`𝒜`$ be an algebra over an arbitrary field $`,`$ and let $`M_n(𝒜)`$ be the $`n\times n`$ total matrix algebra with elements in $`𝒜,`$ and with its basis $`\{\tau _{ij}\}`$ satisfying the multiplication rules $$\tau _{ij}\tau _{st}=\{\begin{array}{c}\tau _{it}j=s,\\ 0js,\end{array}fori,j,s,t=1,2,\mathrm{},n.$$ $`(1.10)`$ Then any $`a=_{i,j=1}^na_{ij}\tau _{ij}M_n(𝒜),`$ where $`a_{ij}𝒜,`$ satisfies the following universal similarity factorization equality $$P\left[\begin{array}{cccc}a& & & \\ & a& & \\ & & \mathrm{}& \\ & & & a\end{array}\right]P^1=\left[\begin{array}{cccc}a_{11}& a_{12}& \mathrm{}& a_{1n}\\ a_{21}& a_{22}& \mathrm{}& a_{2n}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ a_{n1}& a_{n2}& \mathrm{}& a_{nn}\end{array}\right],$$ $`(1.11)`$ where $`P`$ has the following independent form $$P=P^1=\left[\begin{array}{cccc}\tau _{11}& \tau _{21}& \mathrm{}& \tau _{n1}\\ \tau _{12}& \tau _{22}& \mathrm{}& \tau _{n2}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \tau _{1n}& \tau _{2n}& \mathrm{}& \tau _{nn}\end{array}\right].$$ $`(1.12)`$ The correctness of this result can directly be verified by multiplying the left-hand side of Eq.(1.11). The significance of this result is in that if an algebra $``$ is known to be algebraically isomorphic to an $`n\times n`$ total matrix algebra over $`𝒜`$, then there exists an independent invertible matrix $`P`$ over $``$ such that $`P(aI_n)P^1𝒜^{n\times n}`$ holds for all $`a`$. The most part of the results in the paper are established through this basic identity. 2. UNIVERSAL SIMILARITY EQUALITIES OVER $`_{p,q}`$ WITH $`p+q8`$ This section is divided into eight subsections. In the first four of which, we present the universal similarity factorization equalities over $`_{p,q}`$ corresponding to $`p+q4`$ with proofs, and then list the universal similarity factorization equalities over $`_{p,q}`$ corresponding to $`5<p+q8`$. 2.1. The Cases $`_{p,q}`$ with $`p+q=1`$ The two algebras $`_{p,q}`$ corresponding to $`p+q=1`$ are $`_{1,0}`$, the hyperbolic numbers, and $`_{0,1},`$ the complex numbers, respectively. The two fundamental universal similarity factorization equalities over them are very simple but crucial for the subsequent results. Theorem 2.1.1. Let $`a=a_0+a_1e_1_{1,0}`$ be given, where $`a_0,a_1,e_1^2=1,`$ and denote its conjugate $`\overline{a}=a_0a_1e_1`$. Then $`a`$ and $`\overline{a}`$ satisfy the following universal similarity factorization equality $$P_{1,0}\left[\begin{array}{cc}\hfill a& \hfill 0\\ \hfill 0& \hfill \overline{a}\end{array}\right]P_{1,0}^1=\left[\begin{array}{cc}a_0+a_1& 0\\ 0& a_0a_1\end{array}\right]\varphi _{1,0}(a)^2,$$ $`(\mathrm{2.1.1})`$ where $`P_{1,0}`$ and $`P_{1,0}^1`$ have the independent forms $`(`$no relation with $`a)`$ $$P_{1,0}=\frac{1}{2}\left[\begin{array}{cc}1+e_1& (1e_1)\\ 1e_1& 1+e_1\end{array}\right],P_{1,0}^1=\frac{1}{2}\left[\begin{array}{cc}1+e_1& 1e_1\\ (1e_1)& 1+e_1\end{array}\right].$$ $`(\mathrm{2.1.2})`$ Proof. Let $`s=1+e_1`$. Then $`\overline{s}=1e_1`$, and both of them satisfy $`s^2=2s,\overline{s}^2=2\overline{s}`$ and $`s\overline{s}=\overline{s}s=0.`$ Thus it follows that $`P_{1,0}\left[\begin{array}{cc}a& 0\\ 0& \overline{a}\end{array}\right]P_{1,0}^1`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left[\begin{array}{cc}s& \hfill \overline{s}\\ \overline{s}& \hfill s\end{array}\right]\left[\begin{array}{cc}a& 0\\ 0& \overline{a}\end{array}\right]\left[\begin{array}{cc}s& \overline{s}\\ \overline{s}& s\end{array}\right]`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left[\begin{array}{cc}sas+\overline{s}\overline{a}\overline{s}& sa\overline{s}\overline{s}\overline{a}s\\ \overline{s}as& \overline{s}a\overline{s}+s\overline{a}s\end{array}\right]`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left[\begin{array}{cc}as^2+\overline{a}\overline{s}^2& 0\\ 0& a\overline{s}^2+\overline{a}s^2\end{array}\right]={\displaystyle \frac{1}{2}}\left[\begin{array}{cc}as+\overline{a}\overline{s}& 0\\ 0& a\overline{s}+\overline{a}s\end{array}\right],`$ where the two nonzero terms in it are $$as+\overline{a}\overline{s}=(a_0+a_1e_1)(1+e_1)+(a_0a_1e_1)(1e_1)=2(a_0+a_1),$$ $$a\overline{s}+\overline{a}s=(a_0+a_1e_1)(1e_1)+(a_0a_1e_1)(1+e_1)=2(a_0a_1).$$ Hence we have Eq.(2.1.1). $`\mathrm{}`$ It is easily seen that $$\overline{\overline{a}}=a,\overline{a+b}=\overline{a}+\overline{b},\overline{ab}=\overline{a}\overline{b},\overline{\lambda a}=\overline{a\lambda }=\lambda \overline{a}$$ hold for all $`a,b_{1,0},\lambda `$. Thus by Eq.(2.1.1) it follows that (a) $`a=b\varphi _{1,0}(a)=\varphi _{1,0}(b).`$ (b) $`\varphi _{1,0}(a+b)=\varphi _{1,0}(a)+\varphi _{1,0}(b),\varphi _{1,0}(ab)=\varphi _{1,0}(a)\varphi _{1,0}(b),\varphi _{1,0}(\lambda a)=\lambda \varphi _{1,0}(a).`$ (c) $`\varphi _{1,0}(1)=I_2.`$ (d) $`a=\frac{1}{4}[\mathrm{\hspace{0.17em}1}+e_1,1e_1]\varphi _{1,0}(a)[\mathrm{\hspace{0.17em}1}+e_1,1e_1]^T.`$ (e) $`\mathrm{det}[\varphi _{1,0}(a)]=a_0^2a_1^2`$ for all $`a=a_0+a_1e_1_{1,0}.`$ (f) $`a`$ is invertible $`\varphi _{1,0}(a)`$ is invertible, in which case, $`\varphi _{1,0}(a^1)=\varphi _{1,0}^1(a).`$ (g) $`p_a(a)=0,`$ where $`p_a(\lambda )=\mathrm{det}[\lambda I_2\varphi _{1,0}(a)].`$ These properties imply that through the bijective map $$\varphi _{1,0}:a_{1,0}\varphi _{1,0}(a)^2,$$ the Clifford algebra $`_{1,0}`$ is algebraically isomorphic to the matrix algebra $`{}_{}{}^{2}`$, and $`\varphi _{1,0}(a)`$ is a matrix representation of $`a_{1,0}`$ in $`{}_{}{}^{2}`$. As to the algebra $`_{0,1}=𝒞`$, the following result is quite easy to verify . Theorem 2.1.2. Let $`a=a_0+a_1\epsilon _1_{0,1}=𝒞,`$ where $`a_0,a_1,\epsilon _1^2=1,`$ and denote $`\overline{a}=a_0a_1\epsilon _1.`$ Then $`a`$ and $`\overline{a}`$ satisfy the following universal similarity factorization equality $$P_{0,1}\left[\begin{array}{cc}a& 0\\ 0& \overline{a}\end{array}\right]P_{0,1}^1=\left[\begin{array}{cc}a_0& \hfill a_1\\ a_1& \hfill a_0\end{array}\right]\varphi _{0,1}(a)^{2\times 2},$$ $`(\mathrm{2.1.3})`$ where $`P_{0,1}`$ has the independent form $$P_{0,1}=P_{0,1}^1=\frac{1}{\sqrt{2}}\left[\begin{array}{cc}1& \hfill \epsilon _1\\ \epsilon _1& \hfill 1\end{array}\right].$$ $`(\mathrm{2.1.4})`$ Through the bijective map $`\varphi _{0,1}:a\varphi _{0,1}(a),`$ the algebra $`_{0,1}`$ is algebraically isomorphic to the matrix algebra $`𝒜=\left\{\left[\begin{array}{cc}a_0& \hfill a_1\\ a_1& \hfill a_0\end{array}\right]\right|a_0,a_1\}.`$ The two equalities in Eqs.(2.1.1) and (2.1.4) can be extended to all matrices over $`_{1,0}`$ and $`_{0,1}`$ as follows. Theorem 2.1.3. Let $`A=A_0+A_1e_1_{1,0}^{m\times n},`$ where $`A_0,A_1^{m\times n}`$ and $`e_1^2=1.`$ Then $`A`$ and its conjugate $`\overline{A}=A_0A_1e_1`$ satisfy the following universal factorization equality $$Q_{2m}\left[\begin{array}{cc}A& O\\ O& \overline{A}\end{array}\right]Q_{2n}^1=\left[\begin{array}{cc}A_0+A_1& O\\ O& A_0A_1\end{array}\right]\mathrm{\Phi }_{1,0}(A)^2^{m\times n},$$ $`(\mathrm{2.1.5})`$ where $$Q_{2m}=\frac{1}{2}\left[\begin{array}{cc}(1+e_1)I_m& (1e_1)I_m\\ (1e_1)I_m& (1+e_1)I_m\end{array}\right],Q_{2n}^1=\frac{1}{2}\left[\begin{array}{cc}(1+e_1)I_n& (1e_1)I_n\\ (1e_1)I_n& (1+e_1)I_n\end{array}\right].$$ In particular, when $`m=n`$, Eq.(2.1.5) becomes a universal similarity factorization equality over $`_{1,0}`$. Theorem 2.1.4. Let $`A=A_0+A_1\epsilon _1_{0,1}^{m\times n}=𝒞^{m\times n},`$ where $`A_0,A_1^{m\times n},\epsilon _1^2=1.`$ Then $`A`$ and its conjugate $`\overline{A}=A_0A_1\epsilon _1`$ satisfy the following universal factorization equality $$K_{2m}\left[\begin{array}{cc}A& 0\\ 0& \overline{A}\end{array}\right]K_{2n}^1=\left[\begin{array}{cc}A_0& \hfill A_1\\ A_1& \hfill A_0\end{array}\right]\mathrm{\Phi }_{0,1}(A)^{2m\times 2n},$$ $`(\mathrm{2.1.6})`$ where $$K_{2t}=K_{2t}^1=\frac{1}{\sqrt{2}}\left[\begin{array}{cc}I_t& \hfill \epsilon _1I_t\\ \epsilon _1I_t& \hfill I_t\end{array}\right],t=m,n.$$ In particular, when $`m=n`$, Eq.(2.1.6) becomes a universal similarity factorization equality over $`_{0,1}=𝒞.`$ 2.2. The Cases for $`_{p,q}`$ with $`p+q=2`$ The three algebraic isomorphisms for $`_{p,q}`$ with $`p+q=2`$ are shown in Eq.(1.6). Based on Lemma 1.1 and Theorem 2.1.1, we can establish universal similarity factorization equalities over them as follows. Theorem 2.2.1. Let $`a_{2,0}=\{e_1,e_2|e_1^2=1,e_2^2=1\}.`$ Then $`a`$ can factor as $`a=a_0+a_1e_1+a_2e_2+a_3e_{12},`$ where $`a_0`$$`a_3,`$ and $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{2,0}\left[\begin{array}{cc}a& 0\\ 0& a\end{array}\right]P_{2,0}^1=\left[\begin{array}{cc}a_0+a_1& a_2+a_3\\ a_2a_3& a_0a_1\end{array}\right]\varphi _{2,0}(a)^{2\times 2},$$ $`(\mathrm{2.2.1})`$ where $`P_{2,0}`$ has the independent form $$P_{2,0}=P_{2,0}^1=\frac{1}{2}\left[\begin{array}{cc}1+e_1& e_2e_{12}\\ e_2+e_{12}& 1e_1\end{array}\right].$$ $`(\mathrm{2.2.2})`$ Proof. By Lemma 1.1, we take the change of basis of $`_{2,0}`$ as follows $$\tau _{11}=\frac{1}{2}(1+e_1),\tau _{12}=\frac{1}{2}(e_2+e_{12}),\tau _{21}=\frac{1}{2}(e_2e_{12}),\tau _{22}=\frac{1}{2}(1e_1).$$ $`(\mathrm{2.2.3})`$ Then it is not difficult to verify that this new basis satisfies the multiplication laws in Eq.(1.10). In that case, every $`a=a_0+a_1e_1+a_2e_2+a_3e_{12}_{2,0}`$ can be rewritten as $$a=(a_0+a_1)\tau _{11}+(a_2+a_3)\tau _{12}+(a_2a_3)\tau _{21}+(a_0a_1)\tau _{22}.$$ $`(\mathrm{2.2.4})`$ Substituting Eqs.(2.2.3) and (2.2.4) into Eqs.(1.11) and (1.12), we directly obtain Eqs.(2.2.1) and (2.2.2). $`\mathrm{}`$ It is easily seen from Eq.(2.2.1) that for all $`a,b_{2,0},\lambda `$, the following operation properties hold (a) $`a=b\varphi _{2,0}(a)=\varphi _{2,0}(b).`$ (b) $`\varphi _{2,0}(a+b)=\varphi _{2,0}(a)+\varphi _{2,0}(b),\varphi _{2,0}(ab)=\varphi _{2,0}(a)\varphi _{2,0}(b),\varphi _{2,0}(\lambda a)=\lambda \varphi _{2,0}(a).`$ (c) $`\varphi _{2,0}(1)=I_2.`$ (d) $`a=\frac{1}{4}[\mathrm{\hspace{0.17em}1}+e_1,e_2e_{12}]\varphi _{2,0}(a)[\mathrm{\hspace{0.17em}1}+e_1,e_2e_{12}]^T.`$ (e) $`\mathrm{det}[\varphi _{2,0}(a)]=a_0^2a_1^2a_2^2+a_3^2,`$ for all $`a=a_0+a_1e_1+a_2e_2+a_3e_{12}_{2,0}.`$ (f) $`a`$ is invertible $`\varphi _{2,0}(a)`$ is invertible, in which case, $`\varphi _{2,0}(a^1)=\varphi _{2,0}^1(a).`$ (g) $`p_a(a)=0,`$ where $`p_a(x)=\mathrm{det}[xI_2\varphi _{2,0}(a)].`$ (h) $`a`$ is similar to $`b`$ over $`_{0,2}`$, i.e., there is an invertible $`x_{0,2}`$ such that $`xax^1=b`$ if and only if $`\varphi _{2,0}(a)`$ and $`\varphi _{2,0}(b)`$ are similar over $``$. These properties show that through the bijective map $`\varphi _{2,0}:a_{2,0}\varphi _{2,0}(a)^{2\times 2},`$ the Clifford algebra $`_{2,0}`$ is algebraically isomorphic to the matrix algebra $`^{2\times 2}`$, and $`\varphi _{2,0}(a)`$ is the matrix representation of $`a`$ in $`^{2\times 2}`$. Theorem 2.2.2. Let $`a_{1,1}=\{e_1,\epsilon _1|e_1^2=1,\epsilon _1^2=1\},`$ the split quaternion algebra. Then $`a`$ can factor as $`a=a_0+a_1e_1+a_2\epsilon _1+a_3e_1\epsilon _1,`$ where $`a_0`$$`a_3,`$ and $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{1,1}\left[\begin{array}{cc}a& 0\\ 0& a\end{array}\right]P_{1,1}^1=\left[\begin{array}{cc}a_0+a_1& (a_2+a_3)\\ a_2a_3& a_0a_1\end{array}\right]\varphi _{1,1}(a)^{2\times 2},$$ $`(\mathrm{2.2.5})`$ where $`P_{1,1}`$ has the independent form $$P_{1,1}=P_{1,1}^1=\frac{1}{2}\left[\begin{array}{cc}1+e_1& \epsilon _1e_1\epsilon _1\\ (\epsilon _1+e_1\epsilon _1)& 1e_1\end{array}\right].$$ $`(\mathrm{2.2.6})`$ Proof. By Lemma 1.1, we take the change of basis of $`_{1,1}`$ as follows $$\tau _{11}=\frac{1}{2}(1+e_1),\tau _{12}=\frac{1}{2}(\epsilon _1+e_1\epsilon _1),\tau _{21}=\frac{1}{2}(\epsilon _1e_1\epsilon _1),\tau _{22}=\frac{1}{2}(1e_1).$$ $`(\mathrm{2.2.7})`$ Then it is not difficult to verify that this new basis satisfies the multiplication laws in (1.10). In that case, every $`a=a_0+a_1e_1+a_2\epsilon _1+a_3e_1\epsilon _1,_{1,1}`$ can be rewritten as $$a=(a_0+a_1)\tau _{11}(a_2+a_3)\tau _{12}+(a_2a_3)\tau _{21}+(a_0a_1)\tau _{22}.$$ $`(\mathrm{2.2.8})`$ Substituting Eqs.(2.2.7) and (2.2.8) into Eqs.(1.11) and (1.12), we directly obtain Eqs.(2.2.5) and (2.2.6). $`\mathrm{}`$ Similarly it is easy to verify that through the bijective map $`\varphi _{1,1}:a_{1,1}\varphi _{1,1}(a)^{2\times 2},`$ the Clifford algebra $`_{1,1}`$ is algebraically isomorphic to the matrix algebra $`^{2\times 2}`$, and $`\varphi _{1,1}(a)`$ is the matrix representation of $`a`$ in $`^{2\times 2}`$. As to the Clifford algebra $`_{0,2}=`$, the ordinary quaternion division algebra, we have the following two results. Theorem 2.2.3. Let $`a=a_0+a_1\epsilon _1+a_2\epsilon _2+a_3\epsilon _{12}=\{\epsilon _1,\epsilon _2|\epsilon _1^2=1,\epsilon _2^2=1\},`$ where $`a_0`$$`a_3.`$ Then $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{0,2}\left[\begin{array}{cc}\hfill a& \hfill 0\\ \hfill 0& \hfill a\end{array}\right]P_{0,2}^1=\left[\begin{array}{cc}a_0+a_1\epsilon _1& (a_2+a_3\epsilon _1)\\ a_2a_3\epsilon _1& a_0a_1\epsilon _1\end{array}\right]\varphi _{0,2}(a)𝒞^{2\times 2},$$ $`(\mathrm{2.2.9})`$ where $`P_{0,2}`$ has the independent form $$P_{0,2}=P_{0,2}^1=\frac{1}{\sqrt{2}}\left[\begin{array}{cc}\hfill 1& \hfill \epsilon _1\\ \hfill \epsilon _2& \hfill \epsilon _{12}\end{array}\right].$$ $`(\mathrm{2.2.10})`$ Proof. Note that $`a\epsilon _1a\epsilon _1=2(a_0+a_1\epsilon _1)`$ and $`a+\epsilon _1a\epsilon _1=2(a_2\epsilon _2+a_3\epsilon _{12}).`$ We find $`P_{0,2}\left[\begin{array}{cc}\hfill a& \hfill 0\\ \hfill 0& \hfill a\end{array}\right]P_{0,2}^1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[\begin{array}{cc}a\epsilon _1a\epsilon _1& a\epsilon _2+\epsilon _1a\epsilon _{12}\\ \epsilon _2a+\epsilon _{12}a\epsilon _1& (\epsilon _2a\epsilon _2+\epsilon _{12}a\epsilon _{12})\end{array}\right]`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[\begin{array}{cc}a\epsilon _1a\epsilon _1& (a+\epsilon _1a\epsilon _1)\epsilon _2\\ \epsilon _2(a+\epsilon _1a\epsilon _1)& \epsilon _2(a\epsilon _1a\epsilon _1)\epsilon _2\end{array}\right]`$ $`=`$ $`\left[\begin{array}{cc}a_0+a_1\epsilon _1& (a_2\epsilon _2+a_3\epsilon _{12})\epsilon _2\\ \epsilon _2(a_2\epsilon _2+a_3\epsilon _{12})& \epsilon _2(a_0+a_1\epsilon _1)\epsilon _2\end{array}\right]`$ $`=`$ $`\left[\begin{array}{cc}a_0+a_1\epsilon _1& (a_2+a_3\epsilon _1)\\ a_2a_3\epsilon _1& a_0a_1\epsilon _1\end{array}\right],`$ which is exactly the desired result. $`\mathrm{}`$ Theorem 2.2.4. Let $`a=a_0+a_1\epsilon _1+a_2\epsilon _2+a_3\epsilon _{12},`$ where $`a_0`$$`a_3`$. Then $`aI_4`$ satisfies the following universal similarity factorization equality $$Q_{0,2}\left[\begin{array}{cccc}a& & & \\ & a& & \\ & & a& \\ & & & a\end{array}\right]Q_{0,2}^1=\left[\begin{array}{cccc}\hfill a_0& \hfill a_1& \hfill a_2& \hfill a_3\\ \hfill a_1& \hfill a_0& \hfill a_3& \hfill a_2\\ \hfill a_2& \hfill a_3& \hfill a_0& \hfill a_1\\ \hfill a_3& \hfill a_2& \hfill a_1& \hfill a_0\end{array}\right]\phi _{0,2}(a)^{4\times 4},$$ $`(\mathrm{2.2.11})`$ where $`Q_{0,2}`$ has the independent form $$Q_{0,2}=Q_{0,2}^1=\frac{1}{2}\left[\begin{array}{cccc}1& \epsilon _1& \epsilon _2& \epsilon _{12}\\ \epsilon _1& 1& \epsilon _{12}& \epsilon _2\\ \epsilon _2& \epsilon _{12}& 1& \epsilon _1\\ \epsilon _{12}& \epsilon _2& \epsilon _1& 1\end{array}\right].$$ $`(\mathrm{2.2.12})`$ Proof. It is easy to verify that $$\frac{1}{2}\left[\begin{array}{cc}\hfill 1& \hfill 1\\ \hfill 1& \hfill 1\end{array}\right]\left[\begin{array}{cc}x_0& x_1\\ x_1& x_0\end{array}\right]\left[\begin{array}{cc}\hfill 1& \hfill 1\\ \hfill 1& \hfill 1\end{array}\right]=\left[\begin{array}{cc}x_0+x_1& 0\\ 0& x_0x_1\end{array}\right]$$ $`(\mathrm{2.2.13})`$ holds for any $`x_0,x_1`$. On the basis of (2.2.13), we further obtain $$\left[\begin{array}{cccc}x_0& x_1& x_2& x_3\\ x_1& x_0& x_3& x_2\\ x_2& x_3& x_0& x_1\\ x_3& x_2& x_1& x_0\end{array}\right]=V\left[\begin{array}{cccc}d_1& & & \\ & d_2& & \\ & & d_3& \\ & & & d_4\end{array}\right]V,$$ $`(\mathrm{2.2.14})`$ where $`x_0,x_1,x_2,x_3,`$ and $$d_1=x_0+x_1+x_2+x_3,d_2=x_0+x_1x_2x_3,$$ $$d_3=x_0x_1+x_2x_3,d_4=x_0x_1x_2+x_3,$$ $$V=V^T=V^1=\frac{1}{2}\left[\begin{array}{cccc}\hfill 1& \hfill 1& \hfill 1& \hfill 1\\ \hfill 1& \hfill 1& \hfill 1& \hfill 1\\ \hfill 1& \hfill 1& \hfill 1& \hfill 1\\ \hfill 1& \hfill 1& \hfill 1& \hfill 1\end{array}\right].$$ Replacing $`x_0`$$`x_3`$ in Eq.(2.2.14) now by $`a_0,a_1\epsilon _1,a_2\epsilon _2`$ and $`a_3\epsilon _{12}`$, respectively, we obtain $$A\left[\begin{array}{cccc}a_0& a_1\epsilon _1& a_2\epsilon _2& a_3\epsilon _{12}\\ a_1\epsilon _1& a_0& a_3\epsilon _{12}& a_2\epsilon _2\\ a_2\epsilon _2& a_3\epsilon _{12}& a_0& a_1\epsilon _1\\ a_3\epsilon _{12}& a_2\epsilon _2& a_1\epsilon _1& a_0\end{array}\right]=V\left[\begin{array}{cccc}d_1& & & \\ & d_2& & \\ & & d_3& \\ & & & d_4\end{array}\right]V,$$ $`(\mathrm{2.2.15})`$ where $$d_1=a_0+a_1\epsilon _1+a_2\epsilon _2+a_3\epsilon _{12}=a,d_2=a_0+a_1\epsilon _1a_2\epsilon _2a_3\epsilon _{12},$$ $$d_3=a_0a_1\epsilon _1+a_2\epsilon _2a_3\epsilon _{12},d_4=a_0a_1\epsilon _1a_2\epsilon _2+a_3\epsilon _{12}.$$ It is easy to verify that $`d_2=\epsilon _1a\epsilon _1^1,d_3=\epsilon _2a\epsilon _2^1`$ and $`d_4=\epsilon _{12}a\epsilon _{12}^1.`$ Thus $$\mathrm{diag}(d_1,d_2,d_3,d_4)=J\mathrm{diag}(a,a,a,a)J^1,$$ $`(\mathrm{2.2.16})`$ where $`J=\mathrm{diag}(\mathrm{\hspace{0.17em}1},\epsilon _1,\epsilon _2,\epsilon _{12}).`$ On the other hand, it is easy to verify that $$J^1AJ=\left[\begin{array}{cccc}a_0& a_1\epsilon _1\epsilon _1& a_2\epsilon _2\epsilon _2& a_3\epsilon _{12}\epsilon _{12}\\ a_1\epsilon _1^1\epsilon _1& a_0\epsilon _1^1\epsilon _1& a_3\epsilon _1^1\epsilon _{12}\epsilon _2& a_2\epsilon _1^1\epsilon _2\epsilon _{12}\\ a_2\epsilon _2^1\epsilon _2& a_3\epsilon _2^1\epsilon _{12}\epsilon _1& a_0\epsilon _2^1\epsilon _2& a_1\epsilon _2^1\epsilon _1\epsilon _{12}\\ a_3\epsilon _{12}^1\epsilon _{12}& a_2\epsilon _{12}^1\epsilon _2\epsilon _1& a_1\epsilon _{12}^1\epsilon _1\epsilon _2& a_0\epsilon _{12}^1\epsilon _{12}\end{array}\right]=\left[\begin{array}{cccc}\hfill a_0& \hfill a_1& \hfill a_2& \hfill a_3\\ \hfill a_1& \hfill a_0& \hfill a_3& \hfill a_2\\ \hfill a_2& \hfill a_3& \hfill a_0& \hfill a_1\\ \hfill a_3& \hfill a_2& \hfill a_1& \hfill a_0\end{array}\right]=\varphi (a).$$ Putting Eqs.(2.2.15) and (2.2.16) in it yields $$\varphi (a)=J^1AJ=J^1V\mathrm{diag}(d_1,d_2,d_3,d_4)VJ=(J^1VJ)\mathrm{diag}(a,a,a,a)(J^1VJ).$$ Let $`P=J^1VJ.`$ Then we have Eqs.(2.2.11) and (2.2.12). $`\mathrm{}`$ Just as the results in Theorems 2.1.3 and 2.1.4, the universal similarity factorization equalities in Theorems 2.2.1—2.2.4 can also be extend to all matrices over $`_{2,0},_{1,1}`$ and $`_{0,2}`$, respectively. We leave them to the reader. 2.3. The Cases for $`_{p,q}`$ with $`p+q=3`$ According to the multiplication laws in Eqs.(1.2) and (1.3) we know that for all algebras $`_{p,q}`$ with $`p+q=n`$ odd, the commutative rule $$ae_1e_2\mathrm{}e_n=e_1e_2\mathrm{}e_na$$ holds for all $`a_{p,q}`$. Based on this simple fact, we can write all elements of $`_{p,q}`$ with $`p+q=3`$ in the form $$a=a_0+a_1e_{[3]},$$ $`(\mathrm{2.3.1})`$ where $`a_0,a_1\{e_1,e_2\},`$ or $`a_0,a_1\{e_1,e_3\},`$ or $`a_0,a_1\{e_2,e_3\}`$. From (2.3.1) we can introduce the conjugate of $`a`$ as follows $$\overline{a}=a_0a_1e_{123}.$$ $`(\mathrm{2.3.2})`$ Then it is easy to verify that for all $`a,b_{p,q}`$ and $`\lambda `$, $$\overline{\overline{a}}=a,\overline{a+b}=\overline{a}+\overline{b},\overline{ab}=\overline{a}\overline{b},\overline{\lambda a}=\overline{a\lambda }=\lambda \overline{a}.$$ In this subsection we combine Eqs.(2.3.1) and (2.3.2) with the results in Subsection 2.2 to establish four universal similarity factorization equalities over $`_{p,q}`$ with $`p+q=3`$. Theorem 2.3.1. Let $`a_{3,0}=\{e_1,e_2,e_3\},`$ and write $`a`$ as $`a=a_0+a_1e_{[3]},`$ where $$a_0,a_1_{2,0}=\{e_1,e_2\},e_{[3]}^2=1.$$ Then $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{3,0}\left[\begin{array}{cc}a& 0\\ 0& a\end{array}\right]P_{3,0}^1=\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)e_{[3]}\varphi _{3,0}(a)𝒞^{2\times 2},$$ $`(\mathrm{2.3.3})`$ where $`\varphi _{2,0}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.2.1) and $`P_{3,0}`$ has the independent form $$P_{3,0}=P_{3,0}^1=P_{2,0}=\frac{1}{2}\left[\begin{array}{cc}1+e_1& e_2e_{12}\\ e_2+e_{12}& 1e_1\end{array}\right].$$ $`(\mathrm{2.3.4})`$ Proof. Writing $`aI_2`$ as $`aI_2=a_0I_2+a_1e_{[3]}I_2`$ and multiplying $`P_{2,0}`$ and $`P_{2,0}^1`$ on its both sides, we obtain $`P_{2,0}(aI_2)P_{2,0}^1`$ $`=`$ $`P_{2,0}(a_0I_2)P_{2,0}^1+P_{2,0}(a_1e_{[3]}I_2)P_{2,0}^1`$ $`=`$ $`P_{2,0}(a_0I_2)P_{2,0}^1+P_{2,0}(a_1I_2)P_{2,0}^1e_{[3]}`$ $`=`$ $`\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)e_{[3]}=\varphi _{3,0}(a).`$ Note that $`\varphi _{2,0}(a_t)^{2\times 2}`$ and $`e_{[3]}^2=1.`$ Thus (2.3.3) follows. $`\mathrm{}`$ Obviously, through the bijective map $`\varphi _{3,0}:a_{3,0}\varphi _{3,0}(a)𝒞^{2\times 2},`$ the Clifford algebra $`_{3,0}`$ is algebraically isomorphic to the matrix algebra $`𝒞^{2\times 2}`$, and $`\varphi _{3,0}(a)`$ is the matrix representation of $`a`$ in $`𝒞^{2\times 2}`$. Theorem 2.3.2. Let $`a_{2,1}=\{e_1,e_2,\epsilon _1\},`$ and write $`a`$ as $`a=a_0+a_1e,`$ where $$a_0,a_1_{1,1}=\{e_1,\epsilon _1\},e=e_1e_2\epsilon _1,e^2=1.$$ Then the diagonal matrix $`D_a=\mathrm{diag}(aI_2,\overline{a}I_2)`$ satisfies the following universal similarity factorization equality $$P_{2,1}D_aP_{2,1}^1=\left[\begin{array}{cc}\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)& O\\ O& \varphi _{1,1}(a_0)\varphi _{1,1}(a_1)\end{array}\right]\varphi _{2,1}(a)^2^{2\times 2},$$ $`(\mathrm{2.3.5})`$ where $`\varphi _{1,1}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.2.5), and $$P_{2,1}=\frac{1}{2}\left[\begin{array}{cc}(1+e)P_{1,1}& (1e)P_{1,1}\\ (1e)P_{1,1}& (1+e)P_{1,1}\end{array}\right],P_{2,1}^1=\frac{1}{2}\left[\begin{array}{cc}P_{1,1}^1(1+e)& P_{1,1}^1(1e)\\ P_{1,1}^1(1e)& P_{1,1}^1(1+e)\end{array}\right],$$ $`(\mathrm{2.3.6})`$ $`P_{1,1}`$ and $`P_{1,1}^1`$ are given by Eq.(2.2.6). Proof. Writing $`aI_2`$ as $`aI_2=a_0I_2+a_1eI_2`$ and multiplying $`P_{1,1}`$ and $`P_{1,1}^1`$ in Eq.(2.2.6) on its both sides, we get $`P_{1,1}(aI_2)P_{1,1}^1`$ $`=`$ $`P_{1,1}(a_0I_2)P_{1,1}^1+P_{1,1}(a_1eI_2)P_{1,1}^1`$ $`=`$ $`P_{1,1}(a_0I_2)P_{1,1}^1+P_{1,1}(a_1I_2)P_{1,1}^1e=\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)e=\psi (a),`$ where $`\varphi _{1,1}(a_t)^{2\times 2}(t=0,1).`$ Here we denote $`\overline{\psi }(a)=\varphi _{1,1}(a_0)\varphi _{1,1}(a_1)e.`$ Then $`\psi (\overline{a})=\overline{\psi }(a)`$. Note that $`e^2=1.`$ Thus by Theorem 2.2.1, we can build a matrix and its inverse as follows $$V=\frac{1}{2}\left[\begin{array}{cc}(1+e)I_2& (1e)I_2\\ (1e)I_2& (1+e)I_2\end{array}\right],V^1=\frac{1}{2}\left[\begin{array}{cc}(1+e)I_2& (1e)I_2\\ (1e)I_2& (1+e)I_2\end{array}\right].$$ Now applying them to diag$`(\psi (a),\psi (\overline{a}))`$ we obtain $`V\left[\begin{array}{cc}\psi (a)& O\\ O& \psi (\overline{a})\end{array}\right]V^1`$ $`=`$ $`V\left[\begin{array}{cc}\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)e& O\\ O& \varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)e\end{array}\right]V^1`$ $`=`$ $`\left[\begin{array}{cc}\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)& O\\ O& \varphi _{1,1}(a_0)\varphi _{1,1}(a_1)\end{array}\right].`$ Finally substituting $`\psi (a)=P_{1,1}(aI_2)P_{1,1}^1`$ and $`\psi (\overline{a})=P_{1,1}(\overline{a}I_2)P_{1,1}^1`$ into the left-hand side of the above equality yields Eq.(2.3.5). $`\mathrm{}`$ Theorem 2.3.3. Let $`a_{1,2}=\{e_1,\epsilon _1,\epsilon _2\},`$ and write $`a`$ as $`a=a_0+a_1e,`$ where $$a_0,a_1_{1,1}=\{e_1,\epsilon _1\},e=e_1\epsilon _1\epsilon _2,e^2=1.$$ Then $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{1,2}\left[\begin{array}{cc}a& 0\\ 0& a\end{array}\right]P_{1,2}^1=\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)e\varphi _{1,2}(a)𝒞^{2\times 2},$$ $`(\mathrm{2.3.7})`$ where $`\varphi _{1,1}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.2.5), and $$P_{1,2}=P_{1,2}^1=P_{1,1}=\frac{1}{2}\left[\begin{array}{cc}1+e_1& \epsilon _1e_1\epsilon _1\\ (\epsilon _1+e_1\epsilon _1)& 1e_1\end{array}\right].$$ $`(\mathrm{2.3.8})`$ Proof. Writing $`aI_2=a_0I_2+a_1eI_2`$ and multiplying $`P_{1,1}`$ and $`P_{1,1}^1`$ on its both sides, we get $`P_{1,1}(aI_2)P_{1,1}^1`$ $`=`$ $`P_{1,1}(a_0I_2)P_{1,1}^1+P_{1,1}(a_1I_2)P_{1,1}^1`$ $`=`$ $`P_{1,1}(a_0I_2)P_{1,1}^1+P_{1,1}(a_1I_2)P_{1,1}^1e=\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)e.`$ Note that $`\varphi _{1,1}(a_t)^{2\times 2}`$ and $`e^2=1.`$ Thus Eq.(2.3.7) follows. $`\mathrm{}`$ Theorem 2.3.4. Let $`a_{0,3}=\{\epsilon _1,\epsilon _2,\epsilon _3\},`$ and write $`a`$ as $`a=a_0+a_1\epsilon _{[3]},`$ where $$a_0,a_1_{0,2}=\{\epsilon _1,\epsilon _2\}=,\epsilon _{[3]}^2=1.$$ Then $`a`$ and $`\overline{a}=a_0a_1\epsilon _{[3]}`$ satisfy the following universal similarity factorization equality $$P_{0,3}\left[\begin{array}{cc}a& 0\\ 0& \overline{a}\end{array}\right]P_{0,3}^1=\left[\begin{array}{cc}a_0+a_1& 0\\ 0& a_0a_1\end{array}\right]\varphi _{0,3}(a)^2,$$ $`(\mathrm{2.3.9})`$ where $$P_{0,3}=\frac{1}{2}\left[\begin{array}{cc}1+\epsilon _{[3]}& (1\epsilon _{[3]})\\ 1\epsilon _{[3]}& 1+\epsilon _{[3]}\end{array}\right],P_{0,3}^1=\frac{1}{2}\left[\begin{array}{cc}1+\epsilon _{[3]}& 1\epsilon _{[3]}\\ (1\epsilon _{[3]})& 1+\epsilon _{[3]}\end{array}\right].$$ $`(\mathrm{2.3.10})`$ The derivation of Eq.(2.3.5) is much analogous to that of Eq.(2.1.1). So we omit it here. 2.4. The Cases for $`_{p,q}`$ with $`p+q=4`$ The five algebraic isomorphisms for $`_{p,q}`$ with $`p+q=4`$ are shown in Eq.(1.7). Without much effort we can extend the results in Subsection 2.2 to these five algebras. Theorem 2.4.1. Let $`a_{4,0}=\{e_1,e_2,e_3,e_4\}.`$ Then $`a`$ can factor as $$a=a_0+a_1e_{123}+a_2e_{124}+a_3e_{34},$$ where $$a_0,a_1,a_2,a_3_{2,0}=\{e_1,e_2\},$$ $$e_{123}^2=1,e_{124}^2=1,e_{34}=e_{123}e_{124}=e_{124}e_{123}.$$ In that case $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{4,0}\left[\begin{array}{cc}a& 0\\ 0& a\end{array}\right]P_{4,0}^1=\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)e_{123}+\varphi _{2,0}(a_2)e_{124}+\varphi _{2,0}(a_3)e_{34}\varphi _{4,0}(a),$$ $`(\mathrm{2.4.1})`$ where $$\varphi _{4,0}(a)^{2\times 2}\{e_{123},e_{124}\}=^{2\times 2},$$ $`(\mathrm{2.4.2})`$ $`\varphi _{2,0}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.2.1), and $`P_{4,0}=P_{4,0}^1=P_{2,0},`$ the matrix in Eq.(2.2.2). Proof. Note that the commutative rules $`be_{[3]}=e_{[3]}b,be_{124}=e_{124}b`$ and $`be_{34}=e_{34}b`$ hold for all $`b_{2,0}=\{e_1,e_2\}.`$ Thus we immediately obtain $`P_{2,0}(aI_2)P_{2,0}^1`$ $`=`$ $`P_{2,0}(a_0I_2)P_{2,0}^1+P_{2,0}(a_1I_2)P_{2,0}^1e_{[3]}+P_{2,0}(a_2I_2)P_{2,0}^1e_{124}+P_{2,0}(a_3I_2)P_{2,0}^1e_{34}`$ $`=`$ $`\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)e_{123}+\varphi _{2,0}(a_2)e_{124}+\varphi _{2,0}(a_3)e_{34},`$ which is exactly the result in Eq.(2.4.1). $`\mathrm{}`$ Theorem 2.4.2. Let $`a_{3,1}=\{e_1,e_2,e_3,\epsilon _1\}.`$ Then $`a`$ can factor as $$a=a_0+a_1(e_{12}\epsilon _1)+a_2e_{[3]}+a_3(e_3\epsilon _1),$$ where $$a_0,a_1,a_2,a_3_{2,0}=\{e_1,e_2\},$$ $$(e_{12}\epsilon _1)^2=1,e_{[3]}^2=1,e_3\epsilon _1=(e_{12}\epsilon _1)e_{[3]}=e_{[3]}(e_{12}\epsilon _1).$$ In that case $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{3,1}(aI_4)P_{3,1}^1=\left[\begin{array}{cc}\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)& [\varphi _{2,0}(a_2)+\varphi _{2,0}(a_3)]\\ \varphi _{2,0}(a_2)\varphi _{2,0}(a_3)& \varphi _{2,0}(a_0)\varphi _{2,0}(a_1)\end{array}\right]\varphi _{3,1}(a)^{4\times 4},$$ $`(\mathrm{2.4.3})`$ where $`\varphi _{2,0}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.2.1) and $$P_{3,1}=P_{3,1}^1=\frac{1}{2}\left[\begin{array}{cc}(1+e_{12}\epsilon _1)P_{2,0}& (e_{[3]}e_3\epsilon _2)P_{2,0}\\ (e_{[3]}e_3\epsilon _1)P_{2,0}& (1+e_{12}\epsilon _1)P_{2,0}\end{array}\right],$$ $`(\mathrm{2.4.4})`$ where $`P_{2,0}`$ is given in Eq.(2.2.2). Proof. Note that the following commutative laws $`be_{[3]}=e_{[3]}b,be_{124}=e_{124}b`$ and $`be_{34}=e_{34}b`$ hold for all $`b_{2,0}=\{e_1,e_2\}.`$ Thus it follows from (2.2.1) that $`P_{2,0}(aI_2)P_{2,0}^1`$ $`=`$ $`P_{2,0}(a_0I_2)P_{2,0}^1+P_{2,0}(a_1I_2)P_{2,0}^1(e_{12}\epsilon _1)+P_{2,0}(a_2I_2)P_{2,0}^1e_{[3]}+P_{2,0}(a_3I_2)P_{2,0}^1(e_3\epsilon _1)`$ $`=`$ $`\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)(e_{12}\epsilon _1)+\varphi _{2,0}(a_2)e_{[3]}+\varphi _{2,0}(a_3)(e_3\epsilon _1)`$ $``$ $`\psi (a)^{2\times 2}\{e_{12}\epsilon _1,e_{[3]}\}.`$ Next building a matrix and its inverse from the basis $`1,e_{12}\epsilon _1,e_{[3]}`$ and $`e_3\epsilon _1`$ as follows $$V=V^1=\frac{1}{2}\left[\begin{array}{cc}(1+e_{12}\epsilon _1)I_2& (e_{[3]}e_3\epsilon _1)I_2\\ (e_{[3]}+e_3\epsilon _1)I_2& (1e_{12}\epsilon _1)I_2\end{array}\right],$$ and applying them to the matrix $`\psi (a)`$ given above, we obtain $$V\left[\begin{array}{cc}\psi (a)& O\\ O& \psi (a)\end{array}\right]V^1=\left[\begin{array}{cc}\varphi _{2,0}(a_0)+\varphi _{2,0}(a_1)& [\varphi _{2,0}(a_2)+\varphi _{2,0}(a_3)]\\ \varphi _{2,0}(a_2)\varphi _{2,0}(a_3)& \varphi _{2,0}(a_0)\varphi _{2,0}(a_1)\end{array}\right]\varphi _{3,1}(a).$$ Finally substituting $`\psi (a)=P_{2,0}(aI_2)P_{2,0}^1`$ into the left-hand side of the above equality yields Eqs.(2.4.3) and (2.4.4). $`\mathrm{}`$ Similarly we have the following. Theorem 2.4.3. Let $`a_{2,2}=\{e_1,e_2,\epsilon _1,\epsilon _2\},`$ the split Clifford algebra. Then $`a`$ can factor as $$a=a_0+a_1(e_{12}\epsilon _1)+a_2(e_1\epsilon _{12})+a_3(\epsilon _2e_2),$$ where $$a_0,a_1,a_2,a_3_{1,1}=\{e_1,\epsilon _1\},$$ $$(e_{12}\epsilon _1)^2=1,(e_1\epsilon _{12})^2=1,\epsilon _2e_2=(e_{12}\epsilon _1)(e_1\epsilon _{12})=(e_1\epsilon _{12})(e_{12}\epsilon _1).$$ In that case $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{2,2}(aI_4)P_{2,2}^1=\left[\begin{array}{cc}\varphi _{1,1}(a_0)+\varphi _{1,1}(a_1)& [\varphi _{1,1}(a_2)+\varphi _{1,1}(a_3)]\\ \varphi _{1,1}(a_2)\varphi _{1,1}(a_3)& \varphi _{1,1}(a_0)\varphi _{1,1}(a_1)\end{array}\right]\varphi _{2,2}(a)^{4\times 4},$$ $`(\mathrm{2.4.5})`$ where $`\varphi _{1,1}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ given in (2.2.1) and $$P_{2,2}=P_{2,2}^1=\frac{1}{2}\left[\begin{array}{cc}(1+e_{12}\epsilon _1)P_{1,1}& (e_1\epsilon _{12}e_2\epsilon _2)P_{1,1}\\ (e_1\epsilon _{12}e_2\epsilon _2)P_{1,1}& (1+e_{12}\epsilon _1)P_{1,1}\end{array}\right],$$ $`(\mathrm{2.4.6})`$ where $`P_{1,1}`$ is given in Eq.(2.2.6). Theorem 2.4.4. Let $`a_{1,3}=\{e_1,\epsilon _1,\epsilon _2,\epsilon _3\}.`$ Then $`a`$ can factor as $$a=a_0+a_1\epsilon _{[3]}+a_2e_1(\epsilon _{12})+a_3(e_1\epsilon _3),$$ where $$a_0,a_1,a_2,a_3_{0,2}=\{\epsilon _1,\epsilon _2\}=,$$ $$\epsilon _{[3]}^2=1,(e_1\epsilon _{12})^2=1,e_1\epsilon _3=\epsilon _{[3]}(e_1\epsilon _{12})=(e_1\epsilon _{12})\epsilon _{[3]}.$$ In that case $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{1,3}(aI_2)P_{1,3}^1=\left[\begin{array}{cc}a_0+a_1& (a_2+a_3)\\ a_2a_3& a_0a_1\end{array}\right]\varphi _{1,3}(a)^{2\times 2},$$ $`(\mathrm{2.4.7})`$ where $$P_{1,3}=P_{1,3}^1=\frac{1}{2}\left[\begin{array}{cc}1+\epsilon _{[3]}& e_1\epsilon _{12}e_1\epsilon _3\\ (e_1\epsilon _{12}e_1\epsilon _3)& 1\epsilon _{[3]}\end{array}\right].$$ $`(\mathrm{2.4.8})`$ Proof. Note that the following commutative laws $`b\epsilon _{[3]}=\epsilon _{[3]}b,b(e_1\epsilon _{12})=(e_1\epsilon _{12})b,`$ $`b(e_1\epsilon _3)=(e_1\epsilon _3)b`$ hold for all $`b_{0,2}=\{\epsilon _1,\epsilon _2\}.`$ Thus by Theorem 2.2.2 we can build the matrix $`P_{1,3}`$ in Eq.(2.4.8) such that $`aI_2`$ satisfies Eq.(2.4.7). $`\mathrm{}`$ Theorem 2.4.5. Let $`a_{0,4}=\{\epsilon _1,\epsilon _2,\epsilon _3,\epsilon _4\}.`$ Then $`a`$ can factor as $$a=a_0+a_1\epsilon _{123}+a_2\epsilon _{124}+a_3\epsilon _{43},$$ where $$a_0,a_1,a_2,a_3_{0,2}=\{\epsilon _1,\epsilon _2\}=,$$ $$\epsilon _{123}^2=1,\epsilon _{124}^2=1,\epsilon _{43}=\epsilon _{123}\epsilon _{124}=\epsilon _{124}\epsilon _{123}.$$ In that case $`aI_2`$ satisfies the following universal similarity factorization equality $$P_{0,4}(aI_2)P_{0,4}^1=\left[\begin{array}{cc}a_0+a_1& a_2+a_3\\ a_2a_3& a_0a_1\end{array}\right]\varphi _{0,4}(a)^{2\times 2},$$ $`(\mathrm{2.4.9})`$ where $$P_{0,4}=P_{0,4}^1=\frac{1}{2}\left[\begin{array}{cc}1+\epsilon _{123}& \epsilon _{124}\epsilon _{43}\\ \epsilon _{124}\epsilon _{34}& 1\epsilon _{123}\end{array}\right].$$ $`(\mathrm{2.4.10})`$ Proof. Follows from the fact $`b\epsilon _{123}=\epsilon _{123}b,b\epsilon _{124}=e_1\epsilon _{124}b,b\epsilon _{43}=\epsilon _{43}b`$ for all $`b_{0,2}=\{\epsilon _1,\epsilon _2\}`$ and Theorem 2.2.1. $`\mathrm{}`$ 2.5. The Cases for $`_{p,q}`$ with $`p+q=5`$ Just as for $`_{p,q}`$ with $`p+q=3`$, we can write all elements of $`_{p,q}`$ with $`p+q=5`$ in the following form $$a=a_0+a_1e_{[5]}=a_0+e_{[5]}a_1,$$ where $`a_0,a_1`$ are the elements of the Clifford algebras defined on any four generators of $`e_1,e_2,\mathrm{},e_5`$. Besides we can also introduce the conjugate of $`a`$ as follows $$\overline{a}=a_0a_1e_{[5]}.$$ Then it is easy to verify that for all $`a,b_{p,q}`$ and $`\lambda `$, $$\overline{\overline{a}}=a,\overline{a+b}=\overline{a}+\overline{b},\overline{ab}=\overline{a}\overline{b},\overline{\lambda a}=\overline{a\lambda }=\lambda \overline{a}.$$ According to the table in Eq.(1.4), we know that the six Clifford algebras $`_{p,q}`$ with $`p+q=5`$ satisfy the following algebraic isomorphisms $$_{5,0}^2^{2\times 2},_{4,1}𝒞^{4\times 4},_{3,2}^2^{4\times 4},$$ $`(\mathrm{2.5.1})`$ $$_{2,3}𝒞^{4\times 4},_{1,4}^2^{2\times 2},_{0,5}𝒞^{4\times 4}.$$ $`(\mathrm{2.5.2})`$ Based on the results in previous subsection we can establish six universal similarity factorization equalities between elements of $`_{p,q}`$ with $`p+q=5`$ and matrices of the six matrix algebras in Eqs.(2.5.1) and (2.5.2). Theorem 2.5.1. Let $`a_{5,0}=\{e_1,\mathrm{},e_5\},`$ and write $`a`$ as $`a=a_0+a_1e_{[5]},`$ where $$a_0,a_1_{4,0}=\{e_1,e_2,e_3,e_4\},e_{[5]}^2=1.$$ Then the diagonal matrix $`D_a=\mathrm{diag}(aI_2,\overline{a}I_2)`$ satisfies the following universal similarity factorization equality $$P_{5,0}D_aP_{5,0}^1=\left[\begin{array}{cc}\varphi _{4,0}(a_0)+\varphi _{4,0}(a_1)& O\\ O& \varphi _{4,0}(a_0)\varphi _{4,0}(a_1)\end{array}\right]\varphi _{5,0}(a)^2^{2\times 2},$$ $`(\mathrm{2.5.3})`$ where $`\varphi _{4,0}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in (2.4.1) and $$P_{5,0}=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[5]})P_{4,0}& (1e_{[5]})P_{4,0}\\ (1e_{[5]})P_{4,0}& (1+e_{[5]})P_{4,0}\end{array}\right],$$ $`(\mathrm{2.5.4})`$ $$P_{5,0}^1=\frac{1}{2}\left[\begin{array}{cc}P_{4,0}^1(1+e_{[5]})& P_{4,0}^1(1e_{[5]})\\ P_{4,0}^1(1e_{[5]})& P_{4,0}^1(1+e_{[5]})\end{array}\right],$$ $`(\mathrm{2.5.5})`$ $`P_{4,0}`$ is given in Theorem 2.4.1. Proof. Note that $`be_{[5]}=e_{[5]}b`$ holds for all $`b_{4,0}`$. By applying (2.4.1) to $`aI_2=a_0I_2+a_1e_{[5]}I_2`$, we get $`P_{4,0}(aI_2)P_{4,0}^1`$ $`=`$ $`P_{4,0}(a_0I_2)P_{4,0}^1+P_{4,0}(a_1I_2)P_{4,0}^1e_{[5]}=\varphi _{4,0}(a_0)+\varphi _{4,0}(a_1)e_{[5]}=\psi (a),`$ and $$P_{4,0}(\overline{a}I_2)P_{4,0}^1=\varphi _{4,0}(a_0)\varphi _{4,0}(a_1)e_{[5]}=\psi (\overline{a}).$$ By Theorem 2.1.1, we construct a matrix and its inverse as follows $$V=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[5]})I_2& (1e_{[5]})I_2\\ (1e_{[5]})I_2& (1+e_{[5]})I_2\end{array}\right],V^1=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[5]})I_2& (1e_{[5]})I_2\\ (1e_{[5]})I_2& (1+e_{[5]})I_2\end{array}\right].$$ Now applying them to diag$`(\psi (a),\psi (\overline{a}))`$ we obtain $`V\left[\begin{array}{cc}\psi (a)& O\\ O& \psi (\overline{a})\end{array}\right]V^1=\left[\begin{array}{cc}\varphi _{4,0}(a_0)+\varphi _{4,0}(a_1)& O\\ O& \varphi _{4,0}(a_0)\varphi _{4,0}(a_1)\end{array}\right].`$ Finally substituting $`\psi (a)=P_{4,0}(aI_2)P_{4,0}^1`$ and $`\psi (\overline{a})=P_{4,0}(\overline{a}I_2)P_{4,0}^1`$ into the left-hand side of the above equality produces Eq.(2.5.3). $`\mathrm{}`$ Theorem 2.5.2. Let $`a_{4,1}=\{e_1,e_2,e_3,e_4,\epsilon _1\},`$ and write $`a`$ as $`a=a_0+a_1(e_{[4]}\epsilon _1),`$ where $$a_0,a_1_{3,1}=\{e_1,e_2,e_3,\epsilon _1\},(e_{[4]}\epsilon _1)^2=1.$$ Then $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{4,1}(aI_4)P_{4,1}^1=\varphi _{3,1}(a_0)+\varphi _{3,1}(a_1)(e_{[4]}\epsilon _1)\varphi _{4,1}(a)𝒞^{4\times 4},$$ $`(\mathrm{2.5.6})`$ where $`\varphi _{3,1}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ given in (2.4.3) and $`P_{4,1}=P_{3,1},`$ the matrix given in (2.4.4). Proof. Follows directly from Eq.(2.4.3). $`\mathrm{}`$ For economizing space, we omit the proofs of all the following several results. Theorem 2.5.3. Let $`a_{3,2}=\{e_1,e_2,e_3,\epsilon _1,\epsilon _2\},`$ and write $`a`$ as $`a=a_0+a_1(e_{[3]}\epsilon _{[2]}),`$ where $$a_0,a_1_{2,2}=\{e_1,e_2,\epsilon _1,\epsilon _2\},(e_{[3]}\epsilon _{[2]})^2=1.$$ $`(\mathrm{2.5.7})`$ Then the diagonal matrix $`D_a=\mathrm{diag}(aI_2,\overline{a}I_2)`$ satisfies the following universal similarity factorization equality $$P_{3,2}D_aP_{3,2}^1=\left[\begin{array}{cc}\varphi _{2,2}(a_0)+\varphi _{2,2}(a_1)& O\\ O& \varphi _{2,2}(a_0)\varphi _{2,2}(a_1)\end{array}\right]\varphi _{3,2}(a)^2^{4\times 4},$$ $`(\mathrm{2.5.8})`$ where $`\varphi _{2,2}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ defined in (2.4.5) and $$P_{3,2}=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[3]}\epsilon _{[2]})P_{2,2}& (1e_{[3]}\epsilon _{[2]})P_{2,2}\\ (1e_{[3]}\epsilon _{[2]})P_{2,2}& (1+e_{[3]}\epsilon _{[2]})P_{2,2}\end{array}\right],$$ $`(\mathrm{2.5.9})`$ $$P_{3,2}^1=\frac{1}{2}\left[\begin{array}{cc}P_{2,2}^1(1+e_{[3]}\epsilon _{[2]})& P_{2,2}^1(1e_{[3]}\epsilon _{[2]})\\ P_{2,2}^1(1e_{[3]}\epsilon _{[2]})& P_{2,2}^1(1+e_{[3]}\epsilon _{[2]})\end{array}\right],$$ $`(\mathrm{2.5.10})`$ $`P_{2,2}`$ and $`P_{2,2}^1`$ are given in Eq.(2.4.6). Theorem 2.5.4. Let $`a_{2,3}=\{e_1,e_2,\epsilon _1,\epsilon _2,\epsilon _3\},`$ and write $`a`$ as $`a=a_0+a_1(e_{[2]}\epsilon _{[3]}),`$ where $$a_0,a_1_{2,2}=\{e_1,e_2,\epsilon _1,\epsilon _2\},(e_{[2]}\epsilon _{[3]})^2=1.$$ Then $`aI_4`$ satisfies the following universal similarity equality $$P_{2,3}(aI_4)P_{2,3}^1=\varphi _{2,2}(a_0)+\varphi _{2,2}(a_1)e_{[2]}\epsilon _{[3]}\varphi _{2,3}(a)𝒞^{4\times 4},$$ $`(\mathrm{2.5.11})`$ where $`\varphi _{2,2}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ defined in Eq.(2.4.5) and $`P_{2,3}=P_{2,2},`$ the matrix given in Eq.(2.4.6). Theorem 2.5.5. Let $`a_{1,4}=\{e_1,\epsilon _1,\epsilon _2,\epsilon _3,\epsilon _4\},`$ and write $`a`$ as $`a=a_0+a_1(e_1\epsilon _{[4]}),`$ where $$a_0,a_1_{1,3}=\{e_1,\epsilon _1,\epsilon _2,\epsilon _3\},(e_1\epsilon _{[4]})^2=1.$$ Then the diagonal matrix $`D_a=\mathrm{diag}(aI_2,\overline{a}I_2)`$ satisfies the following universal similarity factorization equality $$P_{1,4}D_aP_{1,4}^1=\left[\begin{array}{cc}\varphi _{1,3}(a_0)+\varphi _{1,3}(a_1)& O\\ O& \varphi _{1,3}(a_0)\varphi _{1,3}(a_1)\end{array}\right]\varphi _{1,4}(a)^2^{2\times 2},$$ $`(\mathrm{2.5.12})`$ where $`\varphi _{1,3}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.4.7) and $$P_{1,4}=\frac{1}{2}\left[\begin{array}{cc}(1+e_1\epsilon _{[4]})P_{1,3}& (1e_1\epsilon _{[4]})P_{1,3}\\ (1e_1\epsilon _{[4]})P_{1,3}& (1+e_1\epsilon _{[4]})P_{1,3}\end{array}\right],$$ $`(\mathrm{2.5.13})`$ $$P_{1,4}^1=\frac{1}{2}\left[\begin{array}{cc}P_{1,3}^1(1+e_1\epsilon _{[4]})& P_{1,3}^1(1e_1\epsilon _{[4]})\\ P_{1,3}^1(1e_1\epsilon _{[4]})& P_{1,3}^1(1+e_1\epsilon _{[4]})\end{array}\right],$$ $`(\mathrm{2.5.14})`$ $`P_{1,3}`$ and $`P_{1,3}^1`$ are given in Eq.(2.4.8). Theorem 2.5.6. Let $`a_{0,5}=\{\epsilon _1,\mathrm{},\epsilon _5\}.`$ Then $`a`$ can factor as $`a=a_0+a_1\epsilon _{[5]},`$ where $$a_0,a_1_{2,2}=\{\epsilon _{1234},\epsilon _{1235},\epsilon _1,\epsilon _2\},\epsilon _{[5]}^2=1.$$ Then there is an independent invertible matrix $`P_{0,5}`$ over $`_{0,5}`$ such that $`aI_4`$ satisfies the following universal similarity equality $$P_{0,5}(aI_4)P_{0,5}^1=\varphi _{2,2}(a_0)+\varphi _{2,2}(a_1)\epsilon _{[5]}\varphi _{0,5}(a)𝒞^{4\times 4},$$ $`(\mathrm{2.5.15})`$ where $`\varphi _{2,2}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ defined in Eq.(2.4.5). 2.6. The Cases for $`_{p,q}`$ with $`p+q=6`$ According to the table Eq.(1.4), we know that the seven algebraic isomorphisms for the Clifford algebras $`_{p,q}`$ with $`p+q=6`$ are as follows $$_{6,0}^{4\times 4},_{5,1}^{4\times 4},_{4,2}^{8\times 8},_{3,3}^{8\times 8},$$ $`(\mathrm{2.6.1})`$ $$_{2,4}^{4\times 4},_{1,5}^{4\times 4},_{0,6}^{8\times 8}.$$ $`(\mathrm{2.6.2})`$ For economizing space, we omit the proofs of all the results in this subsection. Theorem 2.6.1. Let $`a_{6,0}=\{e_1,\mathrm{},e_6\}.`$ Then $`a`$ can factor as $$a=a_0+a_1(e_{[4]}e_5)+a_2(e_{[4]}e_6)+a_3e_{56}=a_0+(e_{[4]}e_5)a_1+(e_{[4]}e_6)a_2+e_{56}a_3,$$ where $$a_0,a_1,a_2,a_3_{4,0}=\{e_1,e_2,e_3,e_4\}$$ $$(e_{[4]}e_5)^2=1,(e_{[4]}e_6)^2=1,e_{56}=(e_{[4]}e_5)(e_{[4]}e_6)=(e_{[4]}e_6)(e_{[4]}e_5).$$ Moreover, there is an independent invertible matrix $`P_{6,0}`$ over $`_{6,0}`$ such that $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{6,0}(aI_4)P_{6,0}^1=\left[\begin{array}{cc}\varphi _{4,0}(a_0)+\varphi _{4,0}(a_1)& \varphi _{4,0}(a_2)+\varphi _{4,0}(a_3)\\ \varphi _{4,0}(a_2)\varphi _{4,0}(a_3)& \varphi _{4,0}(a_0)\varphi _{4,0}(a_1)\end{array}\right]\varphi _{6,0}(a)^{4\times 4},$$ where $`\varphi _{4,0}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in (2.4.1). Theorem 2.6.2. Let $`a_{5,1}=\{e_1,\mathrm{},e_5,\epsilon _1\}.`$ Then $`a`$ can factor as $$a=a_0+a_1e_{[5]}+a_2(e_{[4]}\epsilon _1)+a_3(e_5\epsilon _1)=a_0+e_{[5]}a_1+(e_{[4]}\epsilon _1)a_2+(e_5\epsilon _1)a_3,$$ where $$a_0,a_1,a_2,a_3_{4,0}=\{e_1,e_2,e_3,e_4\},$$ $$e_{[5]}^2=1,(e_{[4]}\epsilon _1)^2=1,e_5\epsilon _1=e_{[5]}(e_{[4]}\epsilon _1)=(e_{[4]}\epsilon _1)e_{[5]}.$$ Moreover, there is an independent invertible matrix $`P_{5,1}`$ over $`_{5,1}`$ such that $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{5,1}(aI_4)P_{5,1}^1=\left[\begin{array}{cc}\varphi _{4,0}(a_0)+\varphi _{4,0}(a_1)& [\varphi _{4,0}(a_2)+\varphi _{4,0}(a_3)]\\ \varphi _{4,0}(a_2)\varphi _{4,0}(a_3)& \varphi _{4,0}(a_0)\varphi _{4,0}(a_1)\end{array}\right]\varphi _{5,1}(a)^{4\times 4},$$ where $`\varphi _{4,0}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.4.1). Theorem 2.6.3. Let $`a_{4,2}=\{e_1,\mathrm{},e_4,\epsilon _1,\epsilon _2\}.`$ $$a=a_0+a_1(e_{[3]}\epsilon _{[2]})+a_2(e_{[4]}\epsilon _1)+a_3(e_4\epsilon _2)=a_0+(e_{[3]}\epsilon _{[2]})a_1+(e_{[4]}\epsilon _1)a_2+(e_4\epsilon _2)a_3,$$ where $$a_0,a_1,a_2,a_3_{3,1}=\{e_1,e_2,e_3,\epsilon _1\},$$ $$(e_{[3]}\epsilon _{[2]})^2=1,(e_{[4]}\epsilon _1)^2=1,e_4\epsilon _2=(e_{[3]}\epsilon _{[2]})(e_{[4]}\epsilon _1)=(e_{[4]}\epsilon _1)(e_{[3]}\epsilon _{[2]}).$$ Morever, there is an independent invertible matrix $`P_{4,2}`$ over $`_{4,2}`$ such that $`aI_8`$ satisfies the following universal similarity factorization equality $$P_{4,2}(aI_8)P_{4,2}^1=\left[\begin{array}{cc}\varphi _{3,1}(a_0)+\varphi _{3,1}(a_1)& [\varphi _{3,1}(a_2)+\varphi _{3,1}(a_3)]\\ \varphi _{3,1}(a_2)\varphi _{3,1}(a_3)& \varphi _{3,1}(a_0)\varphi _{3,1}(a_1)\end{array}\right]\varphi _{4,2}(a)^{8\times 8},$$ where $`\varphi _{3,1}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ defined in Eq.(2.4.3). Theorem 2.6.4. Let $`a_{3,3}=\{e_1,e_2,e_3,\epsilon _1,\epsilon _2,\epsilon _3\}.`$ Then $`a`$ cn factor as $$a=a_0+a_1(e_{[3]}\epsilon _{[2]})+a_2(e_{[2]}\epsilon _{[3]})+a_3(e_3\epsilon _3)=a_0+(e_{[3]}\epsilon _{[2]})a_1+(e_{[2]}\epsilon _{[3]})a_2+(e_3\epsilon _3)a_3,$$ where $$a_0,a_1,a_2,a_3_{2,2}=\{e_1,e_2,\epsilon _1,\epsilon _2\},$$ $$(e_{[3]}\epsilon _{[2]})^2=1,(e_{[2]}\epsilon _{[3]})^2=1,e_3\epsilon _3=(e_{[3]}\epsilon _{[2]})(e_{[2]}\epsilon _{[3]})=(e_{[2]}\epsilon _{[3]})(e_{[3]}\epsilon _{[2]}).$$ Moreover there is an independent invertible matrix $`P_{3,3}`$ over $`_{3,3}`$ such that $`aI_8`$ satisfies the following universal similarity factorization equality $$P_{3,3}(aI_8)P_{3,3}^1=\left[\begin{array}{cc}\varphi _{2,2}(a_0)+\varphi _{2,2}(a_1)& [\varphi _{2,2}(a_2)+\varphi _{2,2}(a_3)]\\ \varphi _{2,2}(a_2)\varphi _{2,2}(a_3)& \varphi _{2,2}(a_0)\varphi _{2,2}(a_1)\end{array}\right]\varphi _{3,3}(a)^{8\times 8},$$ where $`\varphi _{2,2}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ defined in (2.4.5). Theorem 2.6.5. Let $`a_{2,4}=\{e_1,e_2,\epsilon _1,\epsilon _2,\epsilon _3,\epsilon _4\}.`$ Then $`a`$ can factor as $$a=a_0+a_1(e_1\epsilon _{[4]})+a_2(e_{[2]}\epsilon _{[3]})+a_3(e_4\epsilon _2)=a_0+(e_1\epsilon _{[4]})a_1+(e_{[2]}\epsilon _{[3]})a_2+(e_4\epsilon _2)a_3,$$ where $$a_0,a_1,a_2,a_3_{1,3}=\{e_1,\epsilon _1,\epsilon _2,\epsilon _3\},$$ $$(e_1\epsilon _{[4]})^2=1,(e_{[2]}\epsilon _{[3]})^2=1,e_4\epsilon _2=(e_1\epsilon _{[4]})(e_{[2]}\epsilon _{[3]})=(e_{[2]}\epsilon _{[3]})(e_1\epsilon _{[4]}).$$ Moreover, there is an independent invertible matrix $`P_{2,4}`$ over $`_{2,4}`$ such that $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{2,4}(aI_4)P_{2,4}^1=\left[\begin{array}{cc}\varphi _{1,3}(a_0)+\varphi _{1,3}(a_1)& [\varphi _{1,3}(a_2)+\varphi _{1,3}(a_3)]\\ \varphi _{1,3}(a_2)\varphi _{1,3}(a_3)& \varphi _{1,3}(a_0)\varphi _{1,3}(a_1)\end{array}\right]\varphi _{2,4}(a)^{4\times 4},$$ where $`\varphi _{1,3}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.4.7). Theorem 2.6.6. Let $`a_{1,5}=\{e_1,\epsilon _1,\mathrm{},\epsilon _5\}.`$ Then $`a`$ can factor as $$a=a_0+a_1(e_1\epsilon _{[4]})+a_2\epsilon _{[5]}+a_3(e_1\epsilon _5)=a_0+(e_1\epsilon _{[4]})a_1+\epsilon _{[5]}a_2+(e_1\epsilon _5)a_3,$$ where $$a_0,a_1,a_2,a_3_{0,4}=\{\epsilon _1,\epsilon _2,\epsilon _3,\epsilon _4\},$$ $$(e_1\epsilon _{[4]})^2=1,\epsilon _{[5]}^2=1,e_1\epsilon _5=(e_1\epsilon _{[4]})\epsilon _{[5]}=\epsilon _{[5]}(e_1\epsilon _{[4]}).$$ Moreover, there is an independent invertible matrix $`P_{1,5}`$ over $`_{1,5}`$ such that $`aI_4`$ satisfies the following universal similarity factorization equality $$P_{1,5}(aI_4)P_{1,5}^1=\left[\begin{array}{cc}\varphi _{0,4}(a_0)+\varphi _{0,4}(a_1)& [\varphi _{0,4}(a_2)+\varphi _{0,4}(a_3)]\\ \varphi _{0,4}(a_2)\varphi _{0,6}(a_3)& \varphi _{0,4}(a_0)\varphi _{0,4}(a_1)\end{array}\right]\varphi _{1,5}(a)^{4\times 4},$$ where $`\varphi _{0,4}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{2\times 2}`$ defined in Eq.(2.4.9). Theorem 2.6.7. Let $`a_{0,6}=\{\epsilon _1,\mathrm{},\epsilon _6\}.`$ Then $`a`$ can factor as $$a=a_0+a_1(\epsilon _{[3]}\epsilon _6)+a_2(\epsilon _{[3]}\epsilon _5)+a_3\epsilon _{56}=a_0+(\epsilon _{[3]}\epsilon _6)a_1+(\epsilon _{[3]}\epsilon _5)a_2+\epsilon _{56}a_3,$$ where $$a_0,a_1,a_2,a_3_{3,1}=\{\epsilon _{124},\epsilon _{134},\epsilon _{234},\epsilon _{[3]}\epsilon _{56}\},$$ $$(\epsilon _{[3]}\epsilon _6)^2=1,(\epsilon _{[3]}\epsilon _5)^2=1,\epsilon _{56}=(\epsilon _{[3]}\epsilon _6)(\epsilon _{[3]}\epsilon _5)=(\epsilon _{[3]}\epsilon _5)(\epsilon _{[3]}\epsilon _6).$$ Moreover, there is an independent invertible matrix $`P_{0,6}`$ over $`_{0,6}`$ such that $`aI_8`$ satisfies the following universal similarity equality $$P_{0,6}(aI_8)P_{0,6}^1=\left[\begin{array}{cc}\varphi _{3,1}(a_0)+\varphi _{3,1}(a_1)& \varphi _{3,1}(a_2)+\varphi _{3,1}(a_3)\\ \varphi _{3,1}(a_2)\varphi _{3,1}(a_3)& \varphi _{3,1}(a_0)\varphi _{3,1}(a_1)\end{array}\right]\varphi _{0,6}(a)^{8\times 8},$$ where $`\varphi _{3,1}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{4\times 4}`$ defined in Eq.(2.4.3). 2.7. The Cases for $`_{p,q}`$ with $`p+q=7`$ According to the table (1.4), the eight algebraic isomorphisms for the Clifford algebras $`_{p,q}`$ with $`p+q=7`$ are as follows $$_{7,0}𝒞^{8\times 8},_{6,1}^2^{4\times 4},_{5,2}𝒞^{8\times 8},_{4,3}^2^{8\times 8},$$ $`(\mathrm{2.7.1})`$ $$_{3,4}𝒞^{8\times 8},_{2,5}^2^{4\times 4},_{1,6}𝒞^{8\times 8},_{0,7}^2^{8\times 8}.$$ $`(\mathrm{2.7.2})`$ The derivation of the universal similarity equalities between elements of $`_{p,q}`$ with $`p+q=7`$ and the eight matrix algebras are much analogous to those in Subsection 2.5. Here we only present the results for $`_{7,0}`$ and $`_{0,7}`$. Just as for $`_{p,q}`$ with $`p+q=5`$, we can decompose $`a_{p,q}`$ with $`p+q=7`$ into the following general form $$a=a_0+a_1e_{[7]}=a_0+e_{[7]}a_1,$$ $`(\mathrm{2.7.3})`$ where $`a_0,a_1`$ are elements of the Clifford algebras defined on any six generators of $`e_1,e_2,\mathrm{},e_7`$. and from this decomposition, we introduce the conjugate of $`a`$ as follows $$\overline{a}=a_0a_1e_{[7]}.$$ $`(\mathrm{2.7.4})`$ Then it is easy to verify that for all $`a,b_{p,q},`$ and $`p+q=7,\lambda `$, $$\overline{\overline{a}}=a,\overline{a+b}=\overline{a}+\overline{b},\overline{ab}=\overline{a}\overline{b},\overline{\lambda a}=\overline{a\lambda }=\lambda \overline{a}.$$ $`(\mathrm{2.7.5})`$ Theorem 2.7.1. Let $`a_{7,0}=\{e_1,\mathrm{},e_7\},`$ and write $`a`$ as $`a=a_0+a_1e_{[7]},`$ where $`e_{[7]}^2=1`$, and $$a_0,a_1_{4,2}=\{e_1,e_2,e_3,e_4,e_{[5]}e_6,e_{[5]}e_7\}.$$ Then there is an independent invertible matrix $`P_{7,0}`$ over $`_{7,0}`$ such that $`aI_8`$ satisfies the following universal similarity factorization equality $$P_{7,0}(aI_8)P_{7,0}^1=\varphi _{4,2}(a_0)+\varphi _{4,2}(a_1)e_{[7]}\varphi _{7,0}(a)𝒞^{8\times 8},$$ where $`\varphi _{4,2}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{8\times 8}`$ defined in Theorem 2.6.3. Theorem 2.7.2. Let $`a_{0,7}=\{\epsilon _1,\mathrm{},\epsilon _7\},`$ and write $`a`$ as $`a=a_0+a_1\epsilon _{[7]},`$ where $`\epsilon _{[7]}^2=1`$, and $$a_0,a_1_{0,6}=\{\epsilon _1,\mathrm{},\epsilon _6\}.$$ Then there is an independent invertible matrix $`P_{0,7}`$ over $`_{0,7}`$ such that the diagonal matrix $`D_a=\mathrm{diag}(aI_8,\overline{a}I_8)`$ satisfies the following similarity equality $$P_{0,7}D_aP_{0,7}^1=\left[\begin{array}{cc}\varphi _{0,6}(a_0)+\varphi _{0,6}(a_1)& O\\ O& \varphi _{0,6}(a_0)\varphi _{0,6}(a_1)\end{array}\right]\varphi _{0,7}(a)^2^{8\times 8},$$ where $`\varphi _{0,6}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{8\times 8}`$ defined in Theorem 2.6.7. 2.8. The Cases for $`_{p,q}`$ with $`p+q=8`$ Just as in subsection 2.7, we only examine the two particular cases over $`_{8,0}`$ and $`_{0,8}`$. Theorem 2.8.1. Let $`a_{8,0}=\{e_1,\mathrm{},e_8\}.`$ Then $`a`$ can factor as $$a=a_0+a_1e_{4567}+a_2e_{4568}+a_3e_{78}=a_0+e_{4567}a_1+e_{4568}a_2+e_{78}a_3,$$ where $$a_t_{3,3}=\{e_1,e_2,e_3,e_{[3]}e_{478},e_{[3]}e_{578},e_{[3]}e_{678}|1,1,1,11,1\},$$ $$e_{4567}^2=1,e_{4568}^2=1,e_{78}=(e_{4567})(e_{4568})=(e_{4568})(e_{4567}),$$ $`t=0`$$`3`$. In that case, there is an independent invertible matrix $`P_{8,0}`$ over $`_{8,0}`$ such that $`aI_{16}`$ satisfies the following universal similarity factorization equality $$P_{8,0}(aI_{16})P_{8,0}^1=\left[\begin{array}{cc}\varphi _{3,3}(a_0)+\varphi _{3,3}(a_1)& \varphi _{3,3}(a_2)+\varphi _{3,3}(a_3)\\ \varphi _{3,3}(a_2)\varphi _{3,3}(a_3)& \varphi _{3,3}(a_0)\varphi _{3,3}(a_1)\end{array}\right]\varphi _{8,0}(a)^{16\times 16},$$ where $`\varphi _{3,3}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{8\times 8}`$ defined in Theorem 2.6.4. Theorem 2.8.2. Let $`a_{0,8}=\{\epsilon _1,\mathrm{},\epsilon _8\},`$ Then $`a`$ can factor as $$a=a_0+a_1(\epsilon _{[6]}\epsilon _8)+a_2(\epsilon _{[6]}\epsilon _7)+a_3\epsilon _{78}=a_0+(\epsilon _{[6]}\epsilon _8)a_1+(\epsilon _{[6]}\epsilon _7)a_2+\epsilon _{78}a_3$$ where $$a_0,a_1,a_2,a_3_{0,6}=\{\epsilon _1,\mathrm{},\epsilon _6\},$$ and $$(\epsilon _{[6]}\epsilon _8)^2=1,(\epsilon _{[6]}\epsilon _7)^2=1,\epsilon _{78}=(\epsilon _{[6]}\epsilon _8)(\epsilon _{[6]}\epsilon _7)=(\epsilon _{[6]}\epsilon _7)(\epsilon _{[6]}\epsilon _8).$$ In that case, there is an independent invertible matrix $`P_{0,8}`$ over $`_{0,8}`$ such that $`aI_{16}`$ satisfies the following universal similarity factorization equality $$P_{0,8}(aI_{16})P_{0,8}^1=\left[\begin{array}{cc}\varphi _{0,6}(a_0)+\varphi _{0,6}(a_1)& \varphi _{0,6}(a_2)+\varphi _{0,6}(a_3)\\ \varphi _{0,6}(a_2)\varphi _{0,6}(a_3)& \varphi _{0,6}(a_0)\varphi _{0,6}(a_1)\end{array}\right]\varphi _{0,8}(a)^{16\times 16},$$ where $`\varphi _{0,6}(a_t)(t=0`$$`3)`$ is the matrix representation of $`a_t`$ in $`^{8\times 8}`$ defined in Theorem 2.6.7. 3. UNIVERSAL SIMILARITY EQUALITIES OVER $`_{n+p,n}`$ AND $`_{n,n+q}`$ WITH $`0p,q6`$ On the basis of the results in Section 2, we present in this section several induction formulas for the universal similarity factorization equalities over $`_{n+p,n}`$ and $`_{n,n+q}`$ with $`0p,q6`$. 3.1. The Cases for $`_{n,n}`$ We have shown in Theorems 2.2.2, 2.4.3 and 2.6.4 that for all $`a_{n,n}`$ with $`n=1,2,3`$, the corresponding $`aI_2,aI_4,aI_8`$ are uniformly similar to their real matrix representations, respectively. By induction, we can establish the following general result for the split Clifford algebra $`_{n,n}`$. Theorem 3.1.1. Suppose that there is an independent invertible matrix $`P_{n1,n1}(n1)`$ over $`_{n1,n1}=\{e_1,\mathrm{},e_{n1},\epsilon _1,\mathrm{},\epsilon _{n1}\}`$ such that $$P_{n1,n1}(aI_{2^{n1}})P_{n1,n1}^1\varphi _{n1,n1}(a)^{2^{n1}\times 2^{n1}}$$ $`(\mathrm{3.1.1})`$ holds for all $`a_{n1,n1}`$. Now let $`a_{n,n}=\{e_1,\mathrm{},e_n,\epsilon _1,\mathrm{},\epsilon _n\}.`$ Then it can factor as $`a`$ $`=`$ $`a_0+a_1(e_{[n]}\epsilon _{[n1]})+a_2(e_{[n1]}\epsilon _{[n]})+a_3\mu _{n,n}`$ $`=`$ $`a_0+(e_{[n]}\epsilon _{[n1]})a_1+(e_{[n1]}\epsilon _{[n]})a_2+\mu _{n,n}a_3,`$ where $$a_0,a_1,a_2,a_3_{n1,n1}=\{e_1,\mathrm{},e_{n1},\epsilon _1,\mathrm{},\epsilon _{n1}\},$$ $$(e_{[n]}\epsilon _{[n1]})^2=1,(e_{[n1]}\epsilon _{[n]})^2=1,$$ $$\mu _{n,n}=(e_{[n]}\epsilon _{[n1]})(e_{[n1]}\epsilon _{[n]})=(e_{[n1]}\epsilon _{[n]})(e_{[n]}\epsilon _{[n1]})=(1)^{n1}e_n\epsilon _n.$$ In that case, $`aI_{2^n}`$ satisfies the following universal similarity factorization equality $$\begin{array}{c}\text{ }P_{n,n}(aI_{2^n})P_{n,n}^1\text{ }\hfill \\ \text{ }=\left[\begin{array}{cc}\varphi _{n1,n1}(a_0)+\varphi _{n1,n1}(a_1)& [\varphi _{n1,n1}(a_2)+\varphi _{n1,n1}(a_3)]\\ \varphi _{n1,n1}(a_2)\varphi _{n1,n1}(a_3)& \varphi _{n1,n1}(a_0)\varphi _{n1,n1}(a_1)\end{array}\right]\text{ }\hfill \\ \text{ }\varphi _{n,n}(a)^{2^n\times 2^n},(\mathrm{3.1.2})\text{ }\hfill \end{array}$$ where $$P_{n,n}=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[n]}\epsilon _{[n1]})P_{n1,n1}& (e_{[n1]}\epsilon _{[n]}\mu _{n,n})P_{n1,n1}\\ (e_{[n1]}\epsilon _{[n]}+\mu _{n,n})P_{n1,n1}& (1e_{[n]}\epsilon _{[n1]})P_{n1,n1}\end{array}\right],$$ $`(\mathrm{3.1.3})`$ $$P_{n,n}^1=\frac{1}{2}\left[\begin{array}{cc}P_{n1,n1}^1(1+e_{[n]}\epsilon _{[n1]})& P_{n1,n1}^1(e_{[n1]}\epsilon _{[n]}\mu _{n,n})\\ P_{n1,n1}^1(e_{[n1]}\epsilon _{[n]}+\mu _{n,n})& P_{n1,n1}^1(1e_{[n]}\epsilon _{[n1]})\end{array}\right].$$ $`(\mathrm{3.1.4})`$ Proof. Applying Eq.(3.1.1) to $`aI_{2^{n1}}`$, we obtain $`P_{n1,n1}(aI_{2^{n1}})P_{n1,n1}^1`$ $`=`$ $`\varphi _{n1,n1}(a_0)+\varphi _{n1,n1}(a_1)(e_{[n]}\epsilon _{[n1]})`$ $`+\varphi _{n1,n1}(a_2)(e_{[n1]}\epsilon _{[n]})+\varphi _{n1,n1}(a_3)\mu _{n,n}\psi (a).`$ Next setting $$V=V^1=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[n]}\epsilon _{[n1]})I_{2^{n1}}& (e_{[n1]}\epsilon _{[n]}\mu _{n,n})I_{2^{n1}}\\ (e_{[n1]}\epsilon _{[n]}+\mu _{n,n})I_{2^{n1}}& (1e_{[n]}\epsilon _{[n1]})I_{2^{n1}}\end{array}\right].$$ and applying it to $`D_a=\mathrm{diag}(\psi (a),\psi (a))`$, we find $$VD_aV^1=\left[\begin{array}{cc}\varphi _{n1,n1}(a_0)+\varphi _{n1,n1}(a_1)& [\varphi _{n1,n1}(a_2)+\varphi _{n1,n1}(a_3)]\\ \varphi _{n1,n1}(a_2)\varphi _{n1,n1}(a_3)& \varphi _{n1,n1}(a_0)\varphi _{n1,n1}(a_1)\end{array}\right].$$ Finally substituting $`\psi (a)=P_{n1,n1}(aI_{2^{n1}})P_{n1,n1}^1`$ into its left-hand side yields the desired result. $`\mathrm{}`$ 3.2. The Cases for $`_{n+1,n}`$ For $`_{n+1,n}`$ with $`n=0,1,2,`$ we have the corresponding universal similarity factorization equalities in Theorems 2.1.1, 2.2.2 and 2.3.3. In general, we have the following. Theorem 3.2.1. Let $`a_{n+1,n}=\{e_1,\mathrm{},e_{n+1},\epsilon _1,\mathrm{},\epsilon _n\}`$ be given. Then $`a`$ can factor as $$a=a_0+a_1e=a_0+ea_1,$$ where $$a_0,a_1_{n,n}=\{e_1,\mathrm{},e_n,\epsilon _1,\mathrm{},\epsilon _n\},$$ $$e=e_{[n+1]}\epsilon _{[n]},e^2=1.$$ Moreover define $`\overline{a}=a_0a_1e`$. In that case, $`D_a=\mathrm{diag}(aI_{2^n},\overline{a}I_{2^n})`$ satisfies the following universal similarity factorization equality $$P_{n+1,n}D_aP_{n+1,n}^1=\left[\begin{array}{cc}\varphi _{n,n}(a_0)+\varphi _{n,n}(a_1)& O\\ O& \varphi _{n,n}(a_0)\varphi _{n,n}(a_1)\end{array}\right]\varphi _{n+1,n}(a)^2^{2^n\times 2^n},$$ where $`\varphi _{n,n}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2^n\times 2^n}`$ defined in Eq.(3.1.3) and $$P_{n+1,n}=\frac{1}{2}\left[\begin{array}{cc}(1+e)P_{n,n}& (1e)P_{n,n}\\ (1e)P_{n,n}& (1+e)P_{n,n}\end{array}\right],P_{n+1,n}^1=\frac{1}{2}\left[\begin{array}{cc}P_{n,n}^1(1+e)& P_{n,n}^1(1e)\\ P_{n,n}^1(1e)& P_{n,n}^1(1+e)\end{array}\right],$$ $`P_{n,n}`$ and $`P_{n,n}^1`$ are the matrices in Eqs.(3.1.4) and (3.1.5). Proof. Applying Eq.(3.1.3) to $`aI_{2^n}=a_0I_{2^n}+a_1eI_{2^n}`$ gives $`P_{n,n}(aI_{2^n})P_{n,n}^1`$ $`=`$ $`P_{n,n}(aI_{2^n})P_{n,n}^1+P_{n,n}(a_1I_{2^n})P_{n,n}^1e`$ $`=`$ $`\varphi _{n,n}(a_0)+\varphi _{n,n}(a_1)e\psi (a).`$ Next setting $$V=\frac{1}{2}\left[\begin{array}{cc}(1+e)I_{2^n}& (1e)I_{2^n}\\ (1e)I_{2^n}& (1+e)I_{2^n}\end{array}\right],V^1=\frac{1}{2}\left[\begin{array}{cc}(1+e)I_{2^n}& (1e)I_{2^n}\\ (1e)I_{2^n}& (1+e)I_{2^n}\end{array}\right],$$ and applying them to $`D_a=\mathrm{diag}(\psi (a),\psi (\overline{a}))`$, we get $$V\left[\begin{array}{cc}\psi (a)& O\\ O& \psi (\overline{a})\end{array}\right]V^1=\left[\begin{array}{cc}\varphi _{n,n}(a_0)+\varphi _{n,n}(a_1)& O\\ O& \varphi _{n,n}(a_0)\varphi _{n,n}(a_1)\end{array}\right].$$ Finally substituting $`\psi (a)=P_{n,n}(aI_{2^n})P_{n,n}^1`$ and $`\psi (\overline{a})=P_{n,n}(\overline{a}I_{2^n})P_{n,n}^1`$ into its left-hand side yields the desired result. $`\mathrm{}`$ 3.3. The Cases for $`_{n+2,n}`$ For $`_{n+2,n}`$ with $`n=0,1,2,`$ we have the corresponding universal similarity equalities in Theorems 2.2.1, 2.4.2 and 2.6.3. In general, we have the following. Theorem 3.3.1. Suppose that there is an independent invertible matrix $`P_{n+1,n1}(n1)`$ over $`_{n+1,n1}=\{e_1,\mathrm{},e_{n+1},\epsilon _1,\mathrm{},\epsilon _{n1}\}`$ such that $$P_{n+1,n1}(aI_{2^n})P_{n+1,n1}^1\varphi _{n+1,n1}(a)^{2^n\times 2^n}$$ holds for all $`a_{n+1,n1}`$.Now let $`a_{n+2,n}=\{e_1,\mathrm{},e_{n+2},\epsilon _1,\mathrm{},\epsilon _n\}.`$ Then $`a`$ can factor as $`a`$ $`=`$ $`a_0+a_1(e_{[n+1]}\epsilon _{[n]})+a_2(e_{[n+2]}\epsilon _{[n1]})+a_3\mu _{n+2,n}`$ $`=`$ $`a_0+(e_{[n+1]}\epsilon _{[n]})a_1+(e_{[n+2]}\epsilon _{[n1]})a_2+\mu _{n+2,n}a_3,`$ where $$a_0,a_1,a_2,a_3_{n+1,n1}=\{e_1,\mathrm{},e_{n+1},\epsilon _1,\mathrm{},\epsilon _{n1}\},$$ $$(e_{[n+1]}\epsilon _{[n]})^2=1,(e_{[n+2]}\epsilon _{[n1]})^2=1,$$ $$\mu _{n+2,n}=(e_{[n+1]}\epsilon _{[n]})(e_{[n+2]}\epsilon _{[n1]})=(e_{[n+2]}\epsilon _{[n1]})(e_{[n+1]}\epsilon _{[n]})=(1)^{n+2}e_{n+2}\epsilon _n.$$ In that case, $`aI_{2^{n+1}}`$ satisfies the following universal similarity factorization equality $`P_{n+2,n}(aI_{2^{n+1}})P_{n+2,n}^1`$ $`=`$ $`\left[\begin{array}{cc}\varphi _{n+1,n1}(a_0)+\varphi _{n+1,n1}(a_1)& [\varphi _{n+1,n1}(a_2)+\varphi _{n+1,n1}(a_3)]\\ \varphi _{n+1,n1}(a_2)\varphi _{n+1,n1}(a_3)& \varphi _{n+1,n1}(a_0)\varphi _{n+1,n1}(a_1)\end{array}\right]`$ $``$ $`\varphi _{n+2,n}(a)^{2^{n+1}\times 2^{n+1}},`$ where $$P_{n+2,n}=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[n+1]}\epsilon _{[n]})P_{n+1,n1}& (e_{[n+2]}\epsilon _{[n1]}\mu _{n+2,n})P_{n+1,n1}\\ (e_{[n+2]}\epsilon _{[n1]}+\mu _{n+2,n})P_{n+1,n1}& (1e_{[n+1]}\epsilon _{[n]})P_{n+1,n1}\end{array}\right],$$ $$P_{n+2,n}^1=\frac{1}{2}\left[\begin{array}{cc}P_{n+1,n1}^1(1+e_{[n]}\epsilon _{[n]})& P_{n+1,n1}^1(e_{[n1]}\epsilon _{[n1]}\mu _{n+2,n})\\ P_{n+1,n1}^1(e_{[n+2]}\epsilon _{[n1]}+\mu _{n+2,n})& P_{n+1,n1}^1(1e_{[n+1]}\epsilon _{[n]})\end{array}\right].$$ The proof of this result is much analogous to that of Theorem 3.1.1. so we omit it here, and the proofs of next several results are also omitted . 3.4. The Cases for $`_{n+3,n}`$ For $`_{n+3,n}`$ with $`n=0,1,`$ we have the corresponding universal similarity factorization equalities in Theorems 2.3.1 and 2.5.2. In general, we have the following. Theorem 3.4.1. Let $`a_{n+3,n}=\{e_1,\mathrm{},e_{n+3},\epsilon _1,\mathrm{},\epsilon _n\}.`$ Then $`a`$ can factor as $$a=a_0+a_1e=a_0+ea_1,$$ where $$a_0,a_1_{n+2,n}=\{e_1,\mathrm{},e_{n+2},\epsilon _1,\mathrm{},\epsilon _n\},$$ $$e=e_{[n+3]}\epsilon _{[n]},e^2=1.$$ In that case, $`aI_{2^{n+1}}`$ satisfies the following universal similarity factorization equality $$P_{n+3,n}(aI_{2^{n+1}})P_{n+3,n}^1=\varphi _{n+2,n}(a_0)+\varphi _{n+2,n}(a_1)\varphi _{n+3,n}(a)𝒞^{2^{n+1}\times 2^{n+1}},$$ where $`\varphi _{n+2,n}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2^n\times 2^n}`$ defined in Theorem 3.3.1 and $`P_{n+3,n}=P_{n+2,n},`$ the matrix given in Theorem 3.3.1. 3.5. The Cases for $`_{n+4,n}`$ For $`_{n+4,n}`$ with $`n=0,1,`$ we have the corresponding universal similarity factorization equalities in Theorems 2.4.1 and 2.6.2. In general, we have the following. Theorem 3.5.1. Suppose that there is an independent invertible matrix $`P_{n+3,n1}(n1)`$ over $`_{n+3,n1}=\{e_1,\mathrm{},e_{n+3},\epsilon _1,\mathrm{},\epsilon _{n1}\}`$ such that $$P_{n+3,n1}(aI_{2^n})P_{n+3,n1}^1\varphi _{n+3,n1}(a)^{2^n\times 2^n}$$ holds for all $`a_{n+3,n1}`$. Now let $`a_{n+4,n}=\{e_1,\mathrm{},e_{n+4},\epsilon _1,\mathrm{},\epsilon _n\}.`$ Then $`a`$ can factor as $`a`$ $`=`$ $`a_0+a_1(e_{[n+4]}\epsilon _{[n1]})+a_2(e_{[n+3]}\epsilon _{[n]})+a_3\mu _{n+4,n}`$ $`=`$ $`a_0+(e_{[n+4]}\epsilon _{[n1]})a_1+(e_{[n+3]}\epsilon _{[n]})a_2+\mu _{n+4,n}a_3,`$ where $$a_0,a_1,a_2,a_3_{n+3,n1}=\{e_1,\mathrm{},e_{n+3},\epsilon _1,\mathrm{},\epsilon _{n1}\},$$ $$(e_{[n+4]}\epsilon _{[n1]})^2=1,(e_{[n+3]}\epsilon _{[n]})^2=1,$$ $$\mu _{n+4,n}=(e_{[n+4]}\epsilon _{[n1]})(e_{[n+3]}\epsilon _{[n]})=(e_{[n+3]}\epsilon _{[n]})(e_{[n+4]}\epsilon _{[n1]})=(1)^{n+3}e_{n+4}\epsilon _n.$$ In that case, $`aI_{2^{n+1}}`$ satisfies the following universal similarity factorization equality $`P_{n+2,n}(aI_{2^{n+1}})P_{n+2,n}^1`$ $`=`$ $`\left[\begin{array}{cc}\varphi _{n+3,n1}(a_0)+\varphi _{n+3,n1}(a_1)& [\varphi _{n+3,n1}(a_2)+\varphi _{n+3,n1}(a_3)]\\ \varphi _{n+3,n1}(a_2)\varphi _{n+3,n1}(a_3)& \varphi _{n+3,n1}(a_0)\varphi _{n+3,n1}(a_1)\end{array}\right]`$ $``$ $`\varphi _{n+4,n}(a)^{2^{n+1}\times 2^{n+1}},`$ where $$P_{n+4,n}=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[n+4]}\epsilon _{[n1]})P_{n+3,n1}& (e_{[n+3]}\epsilon _{[n]}\mu _{n+4,n})P_{n+3,n1}\\ (e_{[n+3]}\epsilon _{[n]}+\mu _{n+4,n})P_{n+3,n1}& (1e_{[n+4]}\epsilon _{[n1]})P_{n+3,n1}\end{array}\right],$$ $$P_{n+4,n}^1=\frac{1}{2}\left[\begin{array}{cc}P_{n+3,n1}^1(1+e_{[n+4]}\epsilon _{[n1]})& P_{n+3,n1}^1(e_{[n+3]}\epsilon _{[n]}\mu _{n+4,n})\\ P_{n+3,n1}^1(e_{[n+4]}\epsilon _{[n]}+\mu _{n+4,n})& P_{n+3,n1}^1(1e_{[n+4]}\epsilon _{[n1]})\end{array}\right].$$ 3.6. The Cases for $`_{n+5,n}`$ For $`_{n+5,n}`$ with $`n=0,`$ we have the corresponding universal similarity factorization equality in Theorem 2.5.1. In general, we have the following. Theorem 3.6.1. Let $`a_{n+5,n}=\{e_1,\mathrm{},e_{n+5},\epsilon _1,\mathrm{},\epsilon _n\}.`$ Then $`a`$ can factor as $$a=a_0+a_1e=a_0+ea_1,$$ where $$a_0,a_1_{n+4,n}=\{e_1,\mathrm{},e_{n+4},\epsilon _1,\mathrm{},\epsilon _n\},$$ $$e=e_{[n+5]}\epsilon _{[n]},e^2=(e_{[n+5]}\epsilon _{[n]})^2=1.$$ Moreover define $`\overline{a}=a_0a_1e`$. In that case, $`D_a=\mathrm{diag}(aI_{2^n},\overline{a}I_{2^n})`$ satisfies the following universal similarity factorization equality $`P_{n+5,n}D_aP_{n+5,n}^1`$ $`=`$ $`\left[\begin{array}{cc}\varphi _{n+4,n}(a_0)+\varphi _{n+4,n}(a_1)& O\\ O& \varphi _{n+4,n}(a_0)\varphi _{n+4,n}(a_1)\end{array}\right]`$ $``$ $`\varphi _{n+5,n}(a)^2^{2^{n+1}\times 2^{n+1}},`$ where $`\varphi _{n+4,n}(a_t)(t=0,1)`$ is the matrix representation of $`a_t`$ in $`^{2^n\times 2^n}`$ defined in Theorem 3.5.1 and $$P_{n+5,n}=\frac{1}{2}\left[\begin{array}{cc}(1+e)P_{n+4,n}& (1e)P_{n+4,n}\\ (1e)P_{n+4,n}& (1+e)P_{n+4,n}\end{array}\right],$$ $$P_{n+5,n}^1=\frac{1}{2}\left[\begin{array}{cc}\hfill P_{n+4,n}^1(1+e)& P_{n+4,n}^1(1e)\\ \hfill P_{n+4,n}^1(1e)& P_{n+4,n}^1(1+e)\end{array}\right],$$ $`P_{n+4,n}`$ and $`P_{n+4,n}^1`$ are the matrices given in Theorem 3.5.1. 3.7. The Cases for $`_{n+6,n}`$ For $`_{n+6,n}`$ with $`n=0,`$ we have the corresponding universal similarity factorization equality in Theorem 2.6.1. In general, we have the following. Theorem 3.7.1. Suppose that there is an independent invertible matrix $`P_{n+5,n1}(n1)`$ over $`_{n+5,n1}=\{e_1,\mathrm{},e_{n+5},\epsilon _1,\mathrm{},\epsilon _{n1}\}`$ such that $$P_{n+5,n1}(aI_{2^{n+1}})P_{n+5,n1}^1\varphi _{n+5,n1}(a)^{2^{n+1}\times 2^{n+1}}$$ holds for all $`a_{n+5,n1}`$. Now let $`a_{n+6,n}=\{e_1,\mathrm{},e_{n+6},\epsilon _1,\mathrm{},\epsilon _n\}.`$ Then $`a`$ can factor as $`a`$ $`=`$ $`a_0+a_1(e_{[n+5]}\epsilon _{[n]})+a_2(e_{[n+6]}\epsilon _{[n1]})+a_3\mu _{n+6,n}`$ $`=`$ $`a_0+(e_{[n+5]}\epsilon _{[n]})a_1+(e_{[n+6]}\epsilon _{[n1]})a_2+\mu _{n+6,n}a_3,`$ where $$a_0,a_1,a_2,a_3_{n+5,n1}=\{e_1,\mathrm{},e_{n+5},\epsilon _1,\mathrm{},\epsilon _{n1}\},$$ $$(e_{[n+5]}\epsilon _{[n]})^2=1,(e_{[n+6]}\epsilon _{[n1]})^2=1,$$ $$\mu _{n+6,n}=(e_{[n+5]}\epsilon _{[n]})(e_{[n+6]}\epsilon _{[n1]})=(e_{[n+6]}\epsilon _{[n1]})(e_{[n+5]}\epsilon _{[n]})=(1)^{n+5}e_{n+6}\epsilon _n.$$ In that case, $`aI_{2^{n+2}}`$ satisfies the following universal similarity factorization equality $`P_{n+6,n}(aI_{2^{n+2}})P_{n+6,n}^1`$ $`=`$ $`\left[\begin{array}{cc}\varphi _{n+5,n1}(a_0)+\varphi _{n+5,n1}(a_1)& [\varphi _{n+5,n1}(a_2)+\varphi _{n+5,n1}(a_3)]\\ \varphi _{n+5,n1}(a_2)\varphi _{n+5,n1}(a_3)& \varphi _{n+5,n1}(a_0)\varphi _{n+5,n1}(a_1)\end{array}\right]`$ $``$ $`\varphi _{n+6,n}(a)^{2^{n+2}\times 2^{n+2}},`$ where $$P_{n+6,n}=\frac{1}{2}\left[\begin{array}{cc}(1+e_{[n+5]}\epsilon _{[n]})P_{n+5,n1}& (e_{[n+6]}\epsilon _{[n1]}\mu _{n+6,n})P_{n+5,n1}\\ (e_{[n+6]}\epsilon _{[n1]}+\mu _{n+6,n})P_{n+5,n1}& (1e_{[n+5]}\epsilon _{[n]})P_{n+5,n1}\end{array}\right],$$ $$P_{n+6,n}^1=\frac{1}{2}\left[\begin{array}{cc}P_{n+5,n1}^1(1+e_{[n+5]}\epsilon _{[n]})& P_{n+5,n1}^1(e_{[n+6]}\epsilon _{[n1]}\mu _{n+6,n})\\ P_{n+5,n1}^1(e_{[n+6]}\epsilon _{[n1]}+\mu _{n+6,n})& P_{n+5,n1}^1(1e_{[n+5]}\epsilon _{[n]})\end{array}\right].$$ In the same manner it is not difficult to give the induction formulas for the universal similarity factorization equalities over $`_{n,n+q}`$ with $`1q6,`$ we leave them to the reader. As to $`_{n+7,n}`$ and $`_{n,n+7}`$ with $`n=0`$, the corresponding universal similarity factorization equalities have been given in Subsection 2.7, the general results corresponding to nonzero $`n`$ will be included in the next two sections. 4. UNIVERSAL SIMILARITY EQUALITIES OVER $`_{p+8,0}`$ AND $`_{0,q+8}`$ According to the two formulas in (1.9), we know that $$_{p+8,0}_{p,0}^{16\times 16},_{0,q+8}_{0,q}^{16\times 16}$$ hold for all finite $`p`$ and $`q`$. Now applying the theorems in Subsection 2.8, we have the following two general results on the universal similarity factorization equalities over $`_{p+8,0}`$ and $`_{0,q+8}.`$ Theorem 4.1. Let $$a_{p+8,0}=\{e_1,\mathrm{},e_8,\alpha _1,\mathrm{},\alpha _p|e_i^2=\alpha _j^2=1,i=1,\mathrm{},8,j=1,\mathrm{},p\}.$$ Then $`a`$ can factor as $$a=\underset{A}{}a_A(e_{[8]}\alpha )_A_{8,0}\{e_{[8]}\alpha _1,\mathrm{},e_{[8]}\alpha _p|(e_{[8]}\alpha _j)^2=1,j=1,\mathrm{},p\},$$ where $`A=(j_1,j_2,\mathrm{},j_k)`$ with $`1j_1<j_2<\mathrm{}<j_kp`$ ranging all naturally ordered subsets of $`\{1,2,\mathrm{},p\};`$ $`a_A`$ has the form $$a_A_{8,0}=\{e_1,\mathrm{},e_8\};$$ $`(e_{[8]}\alpha )_A`$ is defined to be $$(e_{[8]}\alpha )_A=(e_{[8]}\alpha )_{(j_1,j_2,\mathrm{},j_k)}(e_{[8]}\alpha _{j_1})(e_{[8]}\alpha _{j_2})\mathrm{}(e_{[8]}\alpha _{j_p}),(e_{[8]}\alpha )_{A=\mathrm{}}e_{[8]};$$ both of which satisfy $$a_A(e_{[8]}\alpha )_A=(e_{[8]}\alpha )_Aa_A.$$ In that case, $`aI_{16}`$ satisfies the following universal similarity factorization equality $$P_{8,0}(aI_{16})P_{8,0}^1=\underset{A}{}\varphi _{8,0}(a_A)(e_{[8]}\alpha )_A\varphi _{p+8,0}(a),$$ $`(4.1)`$ where $`\varphi _{8,0}(a_A)`$ is the matrix representation of $`a_A`$ in $`^{16\times 16}`$ defined in Theorem 2.8.1, $`P_{8,0}`$ is the independent invertible matrix mentioned in Theorem 2.8.1, meanwhile, $$\varphi _{p+8,0}(a)^{16\times 16}\left\{e_{[8]}\alpha _1,\mathrm{},e_{[8]}\alpha _p\right|(e_{[8]}\alpha _j)^2=1,j=1,\mathrm{},p\}=_{p,0}^{16\times 16},$$ $`(4.2)`$ called the matrix representation of $`a_{p+8,0}`$ in $`_{p,0}^{16\times 16}`$. If $`1p8`$ in $`_{p+8,0}`$, applying the results in Section 2 to the $`16\times 16`$ matrix $`\varphi _{p+8,0}(a)`$ in Eq.(4.1) we can establish a universal similarity factorization equality between elements of $`_{p+8,0}`$ and matrices with elements in $``$, or $`𝒞`$, or $`.`$ Similarly we have the following. Theorem 4.2. Let $$a_{0,q+8}=\{\tau _1,\mathrm{},\tau _8,\epsilon _1,\mathrm{},\epsilon _q|\tau _i^2=\epsilon _j^2=1,i=1,\mathrm{},8,j=1,\mathrm{},q\}.$$ Then $`a`$ can factor as $$a=\underset{A}{}a_A(\tau _{[8]}\epsilon )_A_{0,8}\{\tau _{[8]}\epsilon _1,\mathrm{},\tau _{[8]}\epsilon _q|(\tau _{[8]}\epsilon _j)^2=1,j=1,\mathrm{},q\},$$ where $`A=(j_1,j_2,\mathrm{},j_k)`$ with $`1j_1<j_2<\mathrm{}<j_kq`$ ranges all naturally ordered subsets of $`\{1,2,\mathrm{},q\};`$ $`(\tau _{[8]}\epsilon )_A`$ and $`a_A`$ are $$(\tau _{[8]}\epsilon )_A=(\tau _{[8]}\epsilon )_{(j_1,j_2,\mathrm{},j_k)}(\tau _{[8]}\epsilon _{j_1})(\tau _{[8]}\epsilon _{j_2})\mathrm{}(\tau _{[8]}\epsilon _{j_k}),(\tau _{[8]}\epsilon )_{A=\mathrm{}}\tau _{[8]},$$ $$a_A_{0,8}=\{\tau _1,\mathrm{},\tau _8,|\tau _i^2=1,i=1,\mathrm{},8\},$$ and both of them satisfy $$a_A(\tau _{[8]}\epsilon )_A=(\tau _{[8]}\epsilon )_Aa_A.$$ In that case, $`aI_{16}`$ satisfies the following universal similarity factorization equality $$P_{0,8}(aI_{16})P_{0,8}^1=\underset{A}{}\varphi _{0,8}(a_A)(\tau _{[8]}\epsilon )_A\varphi _{0,q+8}(a),$$ $`(4.3)`$ where $`\varphi _{0,8}(a_A)`$ is the matrix representation of $`a_A`$ in $`^{16\times 16}`$ defined in Theorem 2.8.2, $`P_{0,8}`$ is the independent invertible matrix mentioned in Theorem 2.8.2, meanwhile, $$\varphi _{0,q+8}(a)^{16\times 16}\left\{\tau _{[8]}\epsilon _1,\mathrm{},\tau _{[8]}\epsilon _q\right|(\tau _{[8]}\epsilon _j)^2=1,j=1,\mathrm{},q\}=_{0,q}^{16\times 16},$$ $`(4.4)`$ called the matrix representation of $`a_{0,q+8}`$ in $`_{0,q}^{16\times 16}`$. If $`1q8`$ in $`_{0,q+8}`$, then by applying the results in Section 2 to the $`16\times 16`$ matrix $`\varphi _{0,q+8}(a)`$ in Eq.(4.3) we can establish a universal similarity factorization equality between elements of $`_{0,q+8}`$ and matrices over $``$ or $`𝒞`$ or $`.`$ 5. UNIVERSAL SIMILARITY EQUALITIES OVER $`_{p+8,q}`$ AND $`_{p,q+8}`$ For the two algebraic isomorphisms in Eq.(1.8), we have the following two general results. Theorem 5.1. Let $`a_{p+8,q}=\{e_1,\mathrm{},e_8,\alpha _1,\mathrm{},\alpha _p,\epsilon _1,\mathrm{},\epsilon _q|e_i^2=\alpha _j^2=\epsilon _k^2=1,i=1,\mathrm{},8,j=1,\mathrm{},p,k=1,\mathrm{},q\}.`$ Then $`a`$ can factor as $$a=\underset{A}{}a_A(e_{[8]}\alpha \epsilon )_A_{8,0}\{e_{[8]}\alpha _1,\mathrm{},e_{[8]}\alpha _p,e_{[8]}\epsilon _1,\mathrm{},e_{[8]}\epsilon _q\},$$ where $$(e_{[8]}\alpha _j)^2=1,(e_{[8]}\epsilon _k)^2=1,j=1,\mathrm{},p,k=1,\mathrm{},q;$$ $`A=(A_1,A_2)`$ is the combination of the two ordered multiindices $`A_1`$ and $`A_2`$ $$A_1=(j_1,j_2,\mathrm{},j_s),A_2=(k_1,k_2,\mathrm{},k_t),$$ with $`1j_1<j_2<\mathrm{}<j_sp`$ and $`1k_1<k_2<\mathrm{}<k_tq`$ ranging all natually ordered subsets of $`\{1,2,\mathrm{},p\}`$ and $`\{1,2,\mathrm{},q\},`$ respectively; $$(e_{[8]}\alpha \epsilon )_A(e_{[8]}\alpha )_{A_1}(e_{[8]}\epsilon )_{A_2}=(e_{[8]}\alpha _{j_1})(e_{[8]}\alpha _{j_2})\mathrm{}(e_{[8]}\alpha _{j_s})(e_{[8]}\epsilon _{k_1})(e_{[8]}\epsilon _{k_2})\mathrm{}(e_{[8]}\epsilon _{k_t}),$$ $$(e_{[8]}\alpha \epsilon )_{A=\mathrm{}}e_{[8]},$$ $$a_A_{8,0}=\{e_1,\mathrm{},e_8|e_i^2=1,i=1,\mathrm{},8\},$$ with $$a_A(e_{[8]}\alpha \epsilon )_A=(e_{[8]}\alpha \epsilon )_Aa_A$$ always holding. In that case, $`aI_{16}`$ satisfies the following universal similarity factorization equality $$P_{8,0}(aI_{16})P_{8,0}^1=\underset{A}{}\varphi _{8,0}(a_A)(e_{[8]}\alpha \epsilon )_A\varphi _{p+8,q}(a),$$ $`(5.1)`$ where $`\varphi _{8,0}(a_A)`$ is the matrix representation of $`a_A`$ in $`^{16\times 16}`$ defined in Theorem 2.8.1, $`P_{8,0}`$ is the independent invertible matrix mentioned in Theorem 2.8.1, meanwhile, $$\varphi _{p+8,q}(a)^{16\times 16}\{e_{[8]}\alpha _1,\mathrm{},e_{[8]}\alpha _p,e_{[8]}\epsilon _1,\mathrm{},e_{[8]}\epsilon _q\}=_{p,q}^{16\times 16}.$$ $`(5.2)`$ If $`1p8`$ and $`1q8`$ a in $`_{p+8,q}`$, then by applying the results in Section 2 to the $`16\times 16`$ matrix $`\varphi _{p+8,q}(a)`$ in Eq.(5.1) we can establish a universal similarity factorization equality between elements of $`_{p+8,q}`$ and matrices over $``$ or $`𝒞`$ or $`.`$ The result for $`_{p,q+8}`$ is much similar to that of Theorem 5.1, so we omit it here.
warning/0003/quant-ph0003138.html
ar5iv
text
# Spontaneous decay in the presence of dispersing and absorbing bodies: general theory and application to a spherical cavity ## I Introduction It is well known that the spontaneous decay of an excited atom can be strongly modified when it is placed inside a microcavity . There are typically two qualitatively different regimes: the weak-coupling regime and the strong-coupling regime. The weak-coupling regime is characterized by monotonous exponential decay, the decay rate being enhanced or reduced compared to the free-space value depending on whether the atomic transition frequency fits a cavity resonance or not. The strong-coupling regime, in contrast, is characterized by reversible Rabi oscillations where the energy of the initially atom is periodically exchanged between the atom and the field. This usually requires that the emission is in resonance with a high-quality cavity mode. Recent progress in constructing certain types of microcavities such as microspheres has rendered it possible to approach the ultimate quality level determined by intrinsic material losses , so that the question of the influence of absorbing material on spontaneous decay has been of increasing interest. Effects of material losses on the lifetime of an excited atom have been studied within Fermi’s golden-rule approach . In the mode structure of a microsphere without and with absorber dopant atoms, which is modeled by a constant and a Lorentzian dielectric function respectively, is considered. The spontaneous emission rate and the radiation intensity as a function of the atomic transition frequency is examined in for an atom in a Fabry-Perot cavity filled with a Lorentz-type dielectric in the case of strong medium-cavity interaction but weak atom-field interaction. In this paper we present a theory of the spontaneous decay of an excited two-level atom in the presence of arbitrary dispersing and absorbing dielectric bodies. We apply the theory to the spontaneous decay of an atom in a spherical microcavity of given complex-valued refractive-index profile, as it is typically the case in experimental implementations. The formalism enables us to include in the theory absorption and dispersion in a consistent way and to give a unified treatment of spontaneous emission, without restriction to a particular coupling regime. The plan of the paper is as follows. In Section II, a recently developed quantization scheme for the electromagnetic field in the presence of dispersing and absorbing dielectric bodies is extended in order to include in the theory the resonant interaction of the field with a two-level atom. From the Hamiltonian of the composed system, an integral equation governing the temporal evolution of the upper-level-probability amplitude of the atom is derived, the integral kernel being determined by the Green tensor of the classical, phenomenological Maxwell equations for the dielectric-assisted electromagnetic field. General expressions for the emission pattern and the power spectrum are derived in terms of the atomic parameters and, via the Green tensor, the cavity parameters of the dielectric-matter configuration. In Section III, the theory is used to examine the spontaneous decay of an excited two-level atom inside a spherical cavity, with special emphasis on the intrinsic dispersion and absorption of the wall material. Spherical microcavities have been very attractive systems for both fundamental research in cavity quantum electrodynamics and applications in optoelectronics (see, e.g., and references therein). Changes in the level shifts and lifetimes of atoms inside or near the surfaces of spherical micro-structures have been studied theoretically , the latter also experimentally , and results on the strong-coupling regime have been reported . The theoretical results have been typically based on standard mode expansion, which fails for intrinsically dispersing and absorbing material. Here, we consider a spherical three-layered structure, where the middle layer is assumed to be a (single-resonance) band-gap dielectric of Lorentz type, whereas the outer and inner layers are vacuum. The strength of the atom-field coupling, which is essentially determined by the imaginary part of the Green tensor at the position of the atom, is analyzed and the positions, heights, and widths of the possible cavity resonances are calculated. In particular, the most pronounced resonances are observed within the band gap, the widths of which are proportional to the intrinsic damping constant of the medium. In a single-resonance approximation, conditions of the strong-coupling regime and closed expressions for the atomic upper-state-population probability are derived, which are in good agreement with the numerical solutions. Finally, a summary and conclusions are given in Section IV. ## II General formalism ### A Quantization scheme Let us first consider the electromagnetic field in the presence of dispersing and absorbing dielectric bodies without additional atomic sources. Following , we represent the electric-field operator $`\widehat{𝐄}`$ in the form $`\widehat{𝐄}(𝐫)=\widehat{𝐄}^{(+)}(𝐫)+\widehat{𝐄}^{()}(𝐫),\widehat{𝐄}^{()}(𝐫)=\left[\widehat{𝐄}^{(+)}(𝐫)\right]^{}`$ (1) $`\widehat{𝐄}^{(+)}(𝐫)={\displaystyle _0^{\mathrm{}}}𝑑\omega \underset{¯}{\overset{^}{𝐄}}(𝐫,\omega ),`$ (2) and the magnetic-field operator $`\widehat{𝐁}`$ accordingly. The operators $`\underset{¯}{\overset{^}{𝐄}}`$ and $`\underset{¯}{\overset{^}{𝐁}}`$ then satisfy the Maxwell equations $`\mathbf{}\underset{¯}{\overset{^}{𝐁}}(𝐫,\omega )=0,`$ (3) $`\mathbf{}\left[ϵ_0ϵ(𝐫,\omega )\underset{¯}{\overset{^}{𝐄}}(𝐫,\omega )\right]=\underset{¯}{\overset{^}{\rho }}(𝐫,\omega ),`$ (4) $`\mathbf{}\times \underset{¯}{\overset{^}{𝐄}}(𝐫,\omega )=i\omega \underset{¯}{\overset{^}{𝐁}}(𝐫,\omega ),`$ (5) $`\mathbf{}\times \underset{¯}{\overset{^}{𝐁}}(𝐫,\omega )=i{\displaystyle \frac{\omega }{c^2}}ϵ(𝐫,\omega )\underset{¯}{\overset{^}{𝐄}}(𝐫,\omega )+\mu _0\underset{¯}{\overset{^}{𝐣}}(𝐫,\omega ),`$ (6) where the complex permittivity $`ϵ(𝐫,\omega )`$ is a function of frequency and space, the real part ($`ϵ_\mathrm{R}`$) and the imaginary part ($`ϵ_\mathrm{I}`$) of which satisfy (for any $`𝐫`$) the Kramers–Kronig relations. The operator noise charge and current densities $`\underset{¯}{\overset{^}{\rho }}(𝐫,\omega )`$ and $`\underset{¯}{\overset{^}{𝐣}}(𝐫,\omega )`$ respectively, which are associated with absorption, are related to the operator noise polarization $`\underset{¯}{\overset{^}{𝐏}}(𝐫,\omega )`$ as $`\underset{¯}{\overset{^}{\rho }}(𝐫,\omega )=\mathbf{}\underset{¯}{\overset{^}{𝐏}}(𝐫,\omega ),`$ (7) $`\underset{¯}{\overset{^}{𝐣}}(𝐫,\omega )=i\omega \underset{¯}{\overset{^}{𝐏}}(𝐫,\omega ),`$ (8) where $$\underset{¯}{\overset{^}{𝐏}}(𝐫,\omega )=i\sqrt{\frac{\mathrm{}ϵ_0}{\pi }ϵ_\mathrm{I}(𝐫,\omega )}\widehat{𝐟}(𝐫,\omega )$$ (9) with $`[\widehat{f}_i(𝐫,\omega ),\widehat{f}_j^{}(𝐫^{},\omega ^{})]=\delta _{ij}\delta (𝐫𝐫^{})\delta (\omega \omega ^{}),`$ (10) $`[\widehat{f}_i(𝐫,\omega ),\widehat{f}_j(𝐫^{},\omega ^{})]=0=[\widehat{f}_i^{}(𝐫,\omega ),\widehat{f}_j^{}(𝐫^{},\omega ^{})].`$ (11) From Eqs. (3) – (9) it follows that $`\underset{¯}{\overset{^}{𝐄}}`$ can be written in the form $`\underset{¯}{\overset{^}{𝐄}}(𝐫,\omega )=i\sqrt{{\displaystyle \frac{\mathrm{}}{\pi ϵ_0}}}{\displaystyle \frac{\omega ^2}{c^2}}`$ (13) $`\times {\displaystyle }d^3𝐫^{}\sqrt{ϵ_\mathrm{I}(𝐫^{},\omega )}𝑮(𝐫,𝐫^{},\omega )\widehat{𝐟}(𝐫^{},\omega ),`$ and $`\underset{¯}{\overset{^}{𝐁}}`$ $`=`$ $`(i\omega )^1\mathbf{}\times \underset{¯}{\overset{^}{𝐄}}`$ accordingly, where $`𝑮(𝐫,𝐫^{},\omega )`$ is the classical Green tensor satisfying the equation $$[\frac{\omega ^2}{c^2}ϵ(𝐫,\omega )\mathbf{}\times \mathbf{}\times ]𝑮(𝐫,𝐫^{},\omega )=𝜹(𝐫𝐫^{})$$ (14) together with the boundary condition at infinity \[$`𝜹(𝐫)`$ is the dyadic $`\delta `$-function\]. In this way, the electric- and magnetic-field strengths are expressed in terms of a continuum set of bosonic fields $`\widehat{𝐟}`$ and $`\widehat{𝐟}^{}`$, which play the role of the fundamental (dynamical) variables of the composed system (electromagnetic field and the medium including the dissipative system) whose Hamiltonian is $$\widehat{H}=d^3𝐫_0^{\mathrm{}}𝑑\omega \mathrm{}\omega \widehat{𝐟}^{}(𝐫,\omega )\widehat{𝐟}(𝐫,\omega ).$$ (15) Using Eqs. (13) \[together with Eqs. (1) and (2)\], we can also express the scalar potential $`\widehat{\phi }`$ and the vector potential $`\widehat{𝐀}`$ of the electromagnetic field in terms of the fundamental bosonic fields. In particular, in the Coulomb gauge we obtain $`\mathbf{}\widehat{\phi }(𝐫)=\widehat{𝐄}^{}(𝐫),`$ (16) $`\widehat{𝐀}(𝐫)={\displaystyle _0^{\mathrm{}}}𝑑\omega (i\omega )^1\underset{¯}{\overset{^}{𝐄}}^{}(𝐫,\omega )+\mathrm{H}.\mathrm{c}.,`$ (17) where $$\widehat{𝐄}^{()}(𝐫)=d^3𝐫^{}𝜹^{()}(𝐫𝐫^{})\widehat{𝐄}(𝐫^{}),$$ (18) with $`𝜹^{}(𝐫)`$ and $`𝜹^{}(𝐫)`$ being the transverse and longitudinal $`\delta `$-functions respectively. We now consider the interaction of the medium-assisted electromagnetic field with additional point charges $`q_\alpha `$. Applying the minimal-coupling scheme, we may write the complete Hamiltonian in the form $`\widehat{H}={\displaystyle d^3𝐫_0^{\mathrm{}}𝑑\omega \mathrm{}\omega \widehat{𝐟}^{}(𝐫,\omega )\widehat{𝐟}(𝐫,\omega )}`$ (21) $`+{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{1}{2m_\alpha }}\left[\widehat{𝐩}_\alpha q_\alpha \widehat{𝐀}(𝐫_\alpha )\right]\left[\widehat{𝐩}_\alpha q_\alpha \widehat{𝐀}(𝐫_\alpha )\right]`$ $`+{\displaystyle \frac{1}{2}}{\displaystyle d^3𝐫\widehat{\rho }_\mathrm{A}(𝐫)\widehat{\phi }_\mathrm{A}(𝐫)}+{\displaystyle d^3𝐫\widehat{\rho }_\mathrm{A}(𝐫)\widehat{\phi }(𝐫)},`$ where $`\widehat{𝐫}_\alpha `$ is the position operator and $`\widehat{𝐩}_\alpha `$ is the canonical momentum operator of the $`\alpha `$th charged particle of mass $`m_\alpha `$. The Hamiltonian (21) consists of four terms. The first term is the energy (15) observed when the particles are absent. The second term is the kinetic energy of the particles, and the third term is their Coulomb energy, where the potential $`\widehat{\phi }_\mathrm{A}`$ can be given by $$\widehat{\phi }_\mathrm{A}(𝐫)=𝑑𝐫^{}\frac{\widehat{\rho }_\mathrm{A}(𝐫^{})}{4\pi ϵ_0|𝐫𝐫^{}|},$$ (22) with $$\widehat{\rho }_\mathrm{A}(𝐫)=\underset{\alpha }{}q_\alpha \delta (𝐫\widehat{𝐫}_\alpha )$$ (23) being the charge density. The last term is the Coulomb energy of interaction of the particles with the medium. Note that all terms are expressed in terms of the dynamical variables $`\widehat{𝐟}(𝐫,\omega ),\widehat{𝐟}^{}(𝐫,\omega ),\widehat{𝐫}_\alpha ,\widehat{𝐩}_\alpha `$. ### B Dynamics of an excited two-level atom Let us consider a neutral two-level atom (position $`𝐫_\mathrm{A}`$, transition frequency $`\omega _\mathrm{A}`$) that resonantly interacts with radiation via an electric-dipole transition (dipole moment $`𝝁`$). In this case, the electric-dipole approximation and the rotating wave approximation apply, and the minimal-coupling Hamiltonian (21) simplifies to (Appendix A) $`\widehat{H}={\displaystyle d^3𝐫_0^{\mathrm{}}𝑑\omega \mathrm{}\omega \widehat{𝐟}^{}(𝐫,\omega )\widehat{𝐟}(𝐫,\omega )}`$ (25) $`+\frac{1}{2}\mathrm{}\omega _\mathrm{A}\widehat{\sigma }_z[\widehat{\sigma }^{}\widehat{𝐄}^{(+)}(𝐫_\mathrm{A})𝝁+\mathrm{H}.\mathrm{c}.],`$ where $`\widehat{\sigma }`$, $`\widehat{\sigma }^{}`$, and $`\widehat{\sigma }_z`$ are the Pauli operators of the two-level atom. When the atom is initially in the upper state and the rest of the system is in the vacuum, then the system wave function at time $`t`$ can be written as $`|\psi (t)=C_u(t)e^{i(\omega _\mathrm{A}/2)t}|u|\{0\}`$ (27) $`+{\displaystyle d^3𝐫_0^{\mathrm{}}𝑑\omega C_{li}(𝐫,\omega ,t)e^{i(\omega \omega _\mathrm{A}/2)t}|l|\{1_i(𝐫,\omega )\}},`$ where $`|u`$ and $`|l`$ respectively are the upper and lower atomic states, $`|\{0\}`$ is the vacuum state of the rest of the system, and $`|\{1_i(𝐫,\omega )\}`$ is the state, where the latter is excited in a single-quantum Fock state. Here and in the following we adopt the convention of summation over repeated vector-component indices. The Schrödinger equation yields $`\dot{C}_u(t)={\displaystyle \frac{\mu _j}{\sqrt{\pi ϵ_0\mathrm{}}}}{\displaystyle _0^{\mathrm{}}}d\omega {\displaystyle \frac{\omega ^2}{c^2}}{\displaystyle }d^3𝐫[\sqrt{ϵ_\mathrm{I}(𝐫,\omega )}`$ (29) $`\times G_{ji}(𝐫_\mathrm{A},𝐫,\omega )C_{li}(𝐫,\omega ,t)e^{i(\omega \omega _\mathrm{A})t}],`$ $`\dot{C}_{li}(𝐫,\omega ,t)={\displaystyle \frac{\mu _j}{\sqrt{\pi ϵ_0\mathrm{}}}}{\displaystyle \frac{\omega ^2}{c^2}}\sqrt{ϵ_\mathrm{I}(𝐫,\omega )}`$ (31) $`\times G_{ji}^{}(𝐫_\mathrm{A},𝐫,\omega )C_u(t)e^{i(\omega \omega _\mathrm{A})t}.`$ We now substitute the result of formal integration of Eq. (31) \[$`C_{li}(𝐫,\omega ,0)`$ $`=`$ $`0`$\] into Eq. (29). Making use of the relationship $`\mathrm{Im}G_{kl}(𝐫,𝐫^{},\omega )=`$ (33) $`{\displaystyle d^3𝐬\frac{\omega ^2}{c^2}ϵ_\mathrm{I}(𝐬,\omega )G_{km}(𝐫,𝐬,\omega )G_{lm}^{}(𝐫^{},𝐬,\omega )},`$ we obtain the integro-differential equation $$\dot{C}_u(t)=_0^t𝑑t^{}K(tt^{})C_u(t^{}),$$ (34) with the kernel function $`K(tt^{})={\displaystyle \frac{k_\mathrm{A}^2\mu _i\mu _j}{\mathrm{}\pi ϵ_0}}`$ (36) $`\times {\displaystyle _0^{\mathrm{}}}d\omega e^{i(\omega \omega _\mathrm{A})(tt^{})}\mathrm{Im}G_{ij}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )`$ ($`k_\mathrm{A}`$ $`=`$ $`\omega _\mathrm{A}/c`$). In the spirit of the rotating wave approximation used we have set $`\omega `$ $`=`$ $`\omega _\mathrm{A}`$ in the integral in Eq. (36). Taking the time integral of both sides of Eq. (34), we easily derive, on changing the order of integrations on the right-hand side, $$C_u(t)=_0^t𝑑t^{}\overline{K}(tt^{})C_u(t^{})+1$$ (37) \[$`C_u(0)`$ $`=`$ $`1`$\], where $`\overline{K}(tt^{})={\displaystyle \frac{k_\mathrm{A}^2\mu _i\mu _j}{\mathrm{}\pi ϵ_0}}`$ (39) $`\times {\displaystyle _0^{\mathrm{}}}d\omega {\displaystyle \frac{\mathrm{Im}G_{ij}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )}{i(\omega \omega _\mathrm{A})}}[e^{i(\omega \omega _\mathrm{A})(tt^{})}1].`$ The integral equation (37) is a well-known Volterra equation of the second kind. An algorithm for solving such an integral equation numerically can be found, e.g., in . It is worth noting that the integro-differential equation (34) and the equivalent integral equation (37) apply to the spontaneous decay of an atom in the presence of an arbitrary configuration of dispersing and absorbing dielectric bodies. All the matter parameters that are relevant for the atomic evolution are contained, via the Green tensor, in the kernel functions (36) and (39). In particular when absorption is disregarded and the permittivity is regarded as being a real frequency-independent quantity (which of course can change with space), then the formalism yields the results of standard mode decomposition (see, e.g. ). When the Markov approximation applies, i.e., when in a coarse-grained description of the atomic motion memory effects are disregarded, then we may let $$\frac{e^{i(\omega _\mathrm{A}\omega )(tt^{})}1}{i(\omega _\mathrm{A}\omega )}\zeta (\omega _\mathrm{A}\omega )$$ (40) in Eq. (39) \[$`\zeta (x)`$ $`=`$ $`\pi \delta (x)`$ $`+`$ $`i𝒫/x`$; $`𝒫`$ denotes the principal value\], and thus $$\overline{K}(tt^{})=\frac{1}{2}\left(Ai\delta \omega \right),$$ (41) where $$A=\frac{2k_\mathrm{A}^2\mu _i\mu _j}{\mathrm{}ϵ_0}\mathrm{Im}G_{ij}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega _\mathrm{A})$$ (42) and $$\delta \omega =\frac{2k_\mathrm{A}^2\mu _i\mu _j}{\pi \mathrm{}ϵ_0}𝒫_0^{\mathrm{}}𝑑\omega \frac{\mathrm{Im}G_{ij}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )}{\omega \omega _\mathrm{A}}.$$ (43) Substituting into Eq. (37) for the kernel function the expression (41), we obtain the familiar result that $$C_u(t)=\mathrm{exp}\left[\frac{1}{2}(Ai\delta \omega )t\right].$$ (44) Obviously, this result is also obtained if in the integral in Eq. (34) $`C_u(t^{})`$ is replaced by $`C_u(t)`$ and then the integral is approximated by $`\zeta (\omega _A`$ $``$ $`\omega )`$. Note that the expressions (42) and (43) for the decay rate and the line shift, respectively, are in full agreement with the results in . It is well known that the Markov approximation is an excellent approximation for describing the radiative decay of an excited atom in free space. In order to study the case where the atom is surrounded by dielectric matter, we assume that the atom is localized in a more or less small free-space region, so that the Green tensor at the position of the atom reads $$𝑮(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )=𝑮^\mathrm{V}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )+𝑮^\mathrm{R}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega ),$$ (45) where $`𝑮^\mathrm{V}`$ is the vacuum Green tensor, with $$\mathrm{Im}𝑮^\mathrm{V}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )=\frac{\omega }{6\pi c}𝑰$$ (46) (for the vacuum Green tensor, see, e.g., ), and $`𝑮^\mathrm{R}`$ describes the effects of reflections at the (surfaces of discontinuity of the) surrounding medium. The contribution of $`𝑮^\mathrm{V}`$ to $`\overline{K}`$ can then be treated in the Markov approximation. Application of Eqs. (41) – (43) yields the well-known vacuum decay rate $$A_0=\frac{k_\mathrm{A}^3\mu ^2}{3\mathrm{}\pi ϵ_0}$$ (47) and a divergent contribution to the vacuum Lamb shift which may be omitted, since the (renormalized) vacuum Lamb shift may be thought of as being included in the atomic transition frequency $`\omega _\mathrm{A}`$. In this way, Eq. (39) takes the form $`\overline{K}(tt^{})=\frac{1}{2}A_0+{\displaystyle \frac{k_\mathrm{A}^2\mu _i\mu _j}{\mathrm{}\pi ϵ_0}}`$ (49) $`\times {\displaystyle _0^{\mathrm{}}}d\omega {\displaystyle \frac{\mathrm{Im}G_{ij}^\mathrm{R}(𝐫_\mathrm{A},𝐫_\mathrm{A},\omega )}{i(\omega \omega _\mathrm{A})}}[e^{i(\omega \omega _\mathrm{A})(tt^{})}1].`$ The integral equation (37) together with the kernel function (49) can be regarded as the basic equation for studying the influence of an arbitrary configuration of dispersing and absorbing dielectric matter on the spontaneous decay of an excited atom. ### C Emission pattern The intensity of the light registered by a point-like photodetector at position $`𝐫`$ and time $`t`$ is given by $$I(𝐫,t)\psi (t)|\widehat{𝐄}^{()}(𝐫)\widehat{𝐄}^{(+)}(𝐫)|\psi (t).$$ (50) To obtain the emission pattern associated with the spontaneous decay of an excited atom in the presence of dispersing and absorbing dielectric matter, we combine Eqs. (1), (2), (13), and (27). After some algebra we derive, on using Eq. (33), $`I(𝐫,t)={\displaystyle \underset{i}{}}|{\displaystyle \frac{k_\mathrm{A}^2\mu _j}{\pi ϵ_0}}{\displaystyle _0^t}dt^{}[C_u(t^{})`$ (52) $`\times {\displaystyle _0^{\mathrm{}}}d\omega \mathrm{Im}G_{ij}(𝐫,𝐫_\mathrm{A},\omega )e^{i(\omega \omega _\mathrm{A})(tt^{})}]|^2,`$ where we have again set $`\omega `$ $`=`$ $`\omega _\mathrm{A}`$ in the frequency integral. Again, all the relevant dielectric-matter parameters are contained in the Green tensor. In contrast to Eq. (37) together with the kernel function (49), Eq. (52) requires information about the Green tensor at different space points. In particular, its dependence on space and frequency essentially determines the retardation effects. In the simplest case of the atom being in free space we have $`\mu _j\mathrm{Im}G_{ij}^\mathrm{V}(𝐫,𝐫_\mathrm{A},\omega )={\displaystyle \frac{1}{8i\pi \rho }}\left(𝝁{\displaystyle \frac{𝝆𝝆𝝁}{\rho ^2}}\right)_i`$ (54) $`\times \left(e^{i\omega \rho /c}e^{i\omega \rho /c}\right)+𝒪\left(\rho ^2\right)`$ ($`𝝆`$ $`=`$ $`𝐫`$ $``$ $`𝐫_\mathrm{A}`$). We substitute the expressions (44) (with $`A`$ $`=`$ $`A_0`$) and (54) into Eq. (52), calculate the time integral, and extend in the frequency integral the lower limit to $`\mathrm{}`$, which then can be calculated by contour integration, $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\omega \left(e^{i\omega \rho /c}e^{i\omega \rho /c}\right){\displaystyle \frac{e^{(A_0/2+i\omega _\mathrm{A}^{})t}e^{i\omega t}}{i[\omega (\omega _\mathrm{A}^{}iA_0/2)]}}`$ (56) $`=2\pi \mathrm{exp}\left[(\frac{1}{2}A_0i\omega _\mathrm{A}^{})(t\rho /c)\right]\mathrm{\Theta }(t\rho /c),`$ where $`\omega _\mathrm{A}^{}=\omega _\mathrm{A}\frac{1}{2}\delta \omega `$ (57) \[$`\mathrm{\Theta }(x)`$, unit step function\]. We thus derive the well-known (far-field) result that $$I(𝐫,t)=\left(\frac{k_\mathrm{A}^2\mu \mathrm{sin}\theta }{4\pi ϵ_0\rho }\right)^2e^{A_0(t\rho /c)}\mathrm{\Theta }(t\rho /c),$$ (58) where $`\theta `$ is the angle between $`𝝆`$ and $`𝝁`$. Let us return to the general expression (52). If retardation is ignored and the Markov approximation applies, then we can replace, for all $`𝐫`$, $`C_u(t^{})`$ by $`C_u(t)`$ in the time integral in Eq. (52) and approximate the time integral by $`\zeta (\omega _A`$ $``$ $`\omega )`$. We obtain $$I(𝐫,t)=|𝐅(𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})|^2e^{At},$$ (59) where $`F_i(𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})={\displaystyle \frac{k_\mathrm{A}^2\mu _j}{\pi ϵ_0}}`$ (61) $`\times {\displaystyle _0^{\mathrm{}}}d\omega \mathrm{Im}G_{ij}(𝐫,𝐫_\mathrm{A},\omega )\zeta (\omega _\mathrm{A}\omega ).`$ ### D Emitted-light spectrum Next, let us consider the time-dependent power spectrum of the emitted light, which for sufficiently small passband width of the spectral apparatus can be given by (see, e.g., ) $`S(𝐫,\omega _\mathrm{S},T)={\displaystyle _0^T}dt_2{\displaystyle _0^T}dt_1[e^{i\omega _\mathrm{S}(t_2t_1)}`$ (63) $`\times \widehat{𝐄}^{()}(𝐫,t_2)\widehat{𝐄}^{(+)}(𝐫,t_1)],`$ where $`\omega _\mathrm{S}`$ is the setting frequency of the spectral apparatus and $`T`$ is the operating-time interval of the detector. In close analogy to the derivation of Eq. (52), we combine Eqs. (1), (2), (13), and (27) and use the relation (33) to obtain $`S(𝐫,\omega _\mathrm{S},T)={\displaystyle \underset{i}{}}|{\displaystyle \frac{k_\mathrm{A}^2\mu _j}{\pi ϵ_0}}{\displaystyle _0^T}dt_1[e^{i(\omega _\mathrm{S}\omega _\mathrm{A})t_1}{\displaystyle _0^{t_1}}dt^{}C_u(t^{})`$ (65) $`\times {\displaystyle _0^{\mathrm{}}}d\omega \mathrm{Im}G_{ij}(𝐫,𝐫_\mathrm{A},\omega )e^{i(\omega \omega _\mathrm{A})(t_1t^{})}]|^2.`$ Further calculation again requires knowledge of the Green tensor of the problem. Let us apply Eq. (65) to the free-space case. Following the line that has led from Eq. (52) to Eq. (58), we find that $`S(𝐫,\omega _\mathrm{S},T)=\left({\displaystyle \frac{k_\mathrm{A}^2\mu \mathrm{sin}\theta }{4\pi ϵ_0\rho }}\right)^2`$ (67) $`\times \left|{\displaystyle \frac{e^{[A_0/2+i(\omega _\mathrm{S}\omega _\mathrm{A}^{})](T\rho /c)}1}{\omega _\mathrm{S}\omega _\mathrm{A}^{}+iA_0/2}}\right|^2\mathrm{\Theta }(T\rho /c).`$ In particular for $`T`$ $``$ $`\mathrm{}`$, we recognize the well-known Lorentzian: $`\underset{T\mathrm{}}{lim}S(𝐫,\omega _\mathrm{S},T)=\left({\displaystyle \frac{k_\mathrm{A}^2\mu \mathrm{sin}\theta }{4\pi ϵ_0\rho }}\right)^2`$ (69) $`\times {\displaystyle \frac{1}{[\omega _\mathrm{S}(\omega _\mathrm{A}\delta \omega /2)]^2+A_0^2/4}}.`$ When retardation is ignored and the Markov approximation applies, then Eq. (65) can be simplified in a similar way as Eq. (52). In close analogy to the derivation of Eq. (59) we may write $`S(𝐫,\omega _\mathrm{S},T)=|𝐅(𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})|^2`$ (71) $`\times \left|{\displaystyle \frac{e^{\{A/2+i[\omega _\mathrm{S}(\omega _\mathrm{A}\delta \omega /2)]\}T}1}{\omega _\mathrm{S}(\omega _\mathrm{A}\delta \omega /2)+iA/2}}\right|^2,`$ with $`𝐅(𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})`$ being defined by Eq. (61). ## III Application to a spherical cavity ### A The model We apply the formalism developed in Sec. II to the spontaneous decay of an excited two-level atom placed inside a spherical three-layered structure (Fig. 1). The outer layer ($`r`$ $`>`$ $`R_1`$) and the inner layer ($`0`$ $``$ $`r`$ $`<`$ $`R_2`$) are vacuum, while the middle layer ($`R_2`$ $``$ $`r`$ $``$ $`R_1`$) is a dispersing and absorbing dielectric. The Green tensor of the configuration is given in Appendix B. We have performed the calculations assuming a Lorentz-type dielectric with a single resonance (in the relevant frequency region): $$ϵ(\omega )=1+\frac{\omega _\mathrm{P}^2}{\omega _\mathrm{T}^2\omega ^2i\omega \gamma }.$$ (72) Here, $`\omega _\mathrm{P}`$ is the plasma frequency, which is proportional to the square root of the number density of the Lorentz oscillators and plays the role of the coupling constant between the medium polarization and the electromagnetic field, and $`\omega _\mathrm{T}`$ and $`\gamma `$, respectively, are the position and the width of the medium resonance. The plots in Fig. 2 of the real and imaginary parts of the index of refraction $$n(\omega )=\sqrt{ϵ(\omega )}=n_\mathrm{R}(\omega )+in_\mathrm{I}(\omega )$$ (73) illustrate a typical band-gap behavior of the configuration. ### B Weak-coupling regime In the weak-coupling regime, the excited atomic state decays exponentially \[Eq. (44)\]. For simplicity let us assume that the atom is positioned at the center of the cavity. From Eqs. (42), (45), (46), and (B31), the cavity-modified decay rate is then found to be $$A=\overline{A}(\omega _A)A_0,$$ (74) where $`A_0`$ is decay rate in free space, Eq. (47), and $$\overline{A}(\omega )=1+\mathrm{Re}C_N^{33}(\omega ),$$ (75) with $`C_N^{33}(\omega )`$ being given by Eqs. (B15) – (B28). Note that if mode expansion applies, $`\overline{A}(\omega )`$ would correspond to the change of the density of modes due to the presence of the cavity. When far from the medium resonance absorption is disregarded and hence the (frequency-independent) refractive index is assumed to be real, then previous results obtained by mode decomposition can be recovered. In particular, for $`R_2`$ $``$ $`0`$ we recognize the decay rate obtained in for microspheres and liquid droplets. However, it should be pointed out that even far from the medium resonance the imaginary part of the refractive index cannot be set equal to zero in general, since the contribution to the decay rate of the nonradiative decay associated with absorption increases $``$ $`R_2^3`$ for decreasing $`R_2`$ (and nonvanishing imaginary part of the refractive index) . Let us restrict our attention to a true microcavity ($`R_2\omega _\mathrm{A}/c`$ $``$ $`1`$). From Fig. 3 it is seen that the rate of spontaneous decay sensitively depends on the transition frequency. Narrow-band enhancement of spontaneous decay ($`\overline{A}`$ $`>`$ $`1`$) alternates with broadband inhibition ($`\overline{A}`$ $`<`$ $`1`$). The frequencies where the maxima of enhancement are observed correspond to the resonance frequencies of the cavity. Within the band gap the heights and widths of the frequency intervals in which spontaneous decay is feasible are essentially determined by the material losses. Outside the band-gap zone the change of the decay rate is less pronounced because of the relatively large input–output coupling, the (small) material losses being of secondary importance. When $`R_2\omega /c`$ $``$ $`1`$ and $`\mathrm{exp}[n_\mathrm{I}(R_1`$ $``$ $`R_2)\omega /c]`$ $``$ $`1`$, then Eqs. (B15) – (B28) drastically simplify and $`\overline{A}(\omega )`$ \[Eq. (75)\] reads $$\overline{A}(\omega )\mathrm{Re}\left[\frac{n(\omega )i\mathrm{tan}(R_2\omega /c)}{1in(\omega )\mathrm{tan}(R_2\omega /c)}\right].$$ (76) The positions $`\omega _m`$ of the maxima of $`\overline{A}(\omega )`$ can then be obtained from the equation $`d\overline{A}/d\omega `$ $`=`$ $`0`$. As long as $`n(\omega )`$ can be regarded as being slowly varying compared with $`\mathrm{tan}(\omega R_2/c)`$, we may neglect $`dn/d\omega `$ and thus determine the resonance frequencies from the equation $`2n_\mathrm{I}(\omega _m)\mathrm{tan}(R_2\omega _m/c)`$ (78) $`|n(\omega _m)|^21\sqrt{(|n(\omega _m)|^21)^2+4n_\mathrm{I}^2(\omega _m)}.`$ Note that Eq. (78) is exact when it is regarded as conditional equation of $`R_2`$ for a desired resonance frequency. In the band-gap zone we may assume that $`n_\mathrm{I}`$ $``$ $`n_\mathrm{R}`$ (see Fig. 2). From Eqs. (76) and (78) it then follows that the maximum values $`\overline{A}(\omega _m)`$ and half widths at half maximum, $`\delta \omega _m`$, of the cavity resonance lines, i.e., the regions where enhanced spontaneous decay can be observed, are given by $$\overline{A}(\omega _m)\frac{n_\mathrm{I}^2(\omega _m)+1}{n_\mathrm{R}(\omega _m)}\frac{2\sqrt{(\omega _\mathrm{L}^2\omega _m^2)(\omega _m^2\omega _\mathrm{T}^2)}}{\gamma \omega _m},$$ (79) $$\delta \omega _m\frac{c}{R_2\overline{A}(\omega _m)},$$ (80) where we have assumed that the material losses are small, i.e., $`\gamma `$ $``$ $`\omega _\mathrm{T},\omega _\mathrm{P},\omega _\mathrm{P}^2/\omega _\mathrm{T}`$. Equations (79) and (80) reveal that in the approximation made the heights (widths) of the resonance lines (Lorentzians) are proportional (inversely proportional) to $`\gamma `$, the highest and narrowest line being in the center of the band gap. Note that the product $`\overline{A}(\omega _m)\delta \omega _m`$ does not depend on $`\gamma `$. Outside the band-gap zone the inequality $`n_\mathrm{R}`$ $``$ $`n_\mathrm{I}`$ is typically valid (see Fig. 2), and thus Eqs. (76) and (78) yield $$\overline{A}(\omega _m)n_\mathrm{R}(\omega _m)\sqrt{\frac{\omega _\mathrm{L}^2\omega _m^2}{\omega _\mathrm{T}^2\omega _m^2}},$$ (81) $$\delta \omega _m\frac{c}{R_2\overline{A}(\omega _m)}.$$ (82) The heights and widths of the resonance lines are now seen to be (approximately) independent of $`\gamma `$. The lines become higher and narrower if $`\omega _m`$ becomes close to $`\omega _\mathrm{T}`$. The widths of the resonance lines are responsible for the damping of intracavity fields. There are two damping mechanisms: photon leakage to the outside of the cavity and photon absorption by the cavity-wall material. From the analysis given above it is seen that the first mechanism is the dominant one outside the band gap where normal dispersion ($`dn_\mathrm{R}/d\omega `$ $`>`$ $`0`$) is observed, while the latter dominates inside the band gap where anomalous dispersion ($`dn_\mathrm{R}/d\omega `$ $`<`$ $`0`$) is observed. To illustrate this in more detail, we have calculated the amount of radiation energy observed outside the cavity, $$W=2cϵ_0_0^{\mathrm{}}𝑑t_0^{2\pi }𝑑\varphi _0^\pi 𝑑\theta \rho ^2\mathrm{sin}\theta I(𝐫,t)$$ (83) ($`\rho `$ $`>`$ $`R_1`$), and compared it with the emitted energy in free space $`W_0`$ $`=`$ $`\mathrm{}\omega _\mathrm{A}`$. Assuming without loss of generality that the atomic transition dipole is $`z`$-oriented and restricting our attention to the relevant far-field contribution, from Eqs. (59) and (61) together with Eq. (B1) and Eqs. (B35) – (B35) we derive (see Appendix C) $$\frac{W}{W_0}\frac{|A_N^{13}(\omega _\mathrm{A})|^2}{1+\mathrm{Re}C_N^{33}(\omega _\mathrm{A})}.$$ (84) Examples of the dependence on the atomic transition frequency of the amount of radiation energy observed outside the cavity are plotted in Fig. 4. It is seen that inside the gap most of the energy emitted by the atom is absorbed by the cavity wall in the course of time, while outside the gap the absorption is (for the chosen values of $`\gamma `$) much less significant. Note that with increasing value of $`\gamma `$ the band gap is smoothed a little bit, and thus the fraction of light that escapes from the cavity can increase. ### C Strong-coupling regime The strength of the atom-field coupling increases when the atomic transition frequency $`\omega _\mathrm{A}`$ approaches a cavity-resonance frequency $`\omega _m`$. In order to gain insight into the strong-coupling regime, let us first consider the limiting case of one cavity-resonance line being involved in the atom-field interaction. Indeed, when $`\omega _\mathrm{A}`$ is close to $`\omega _m`$, then contributions from the other resonance lines become small. Using Eqs. (42), (74), and (75), and recalling that $`\overline{A}(\omega )`$ behaves like a Lorentzian in the vicinity of $`\omega _m`$, we may simplify Eq. (36) to $`K(tt^{}){\displaystyle \frac{A_0}{2\pi }}\overline{A}(\omega _m)(\delta \omega _m)^2e^{i(\omega _m\omega _\mathrm{A})(tt^{})}`$ (87) $`\times {\displaystyle _{\mathrm{}}^+\mathrm{}}d\omega {\displaystyle \frac{e^{i(\omega \omega _m)(tt^{})}}{(\omega \omega _m)^2+(\delta \omega _m)^2}}`$ $`=\frac{1}{2}A_0\overline{A}(\omega _m)\delta \omega _me^{i(\omega _m\omega _\mathrm{A})(tt^{})}e^{\delta \omega _m|tt^{}|}.`$ Substituting this expression into Eq. (34) and differentiating both sides of the resulting equation with regard to time, we arrive at $$\ddot{C}_u(t)+\left[i(\omega _m\omega _\mathrm{A})+\delta \omega _m\right]\dot{C}_u(t)+(\mathrm{\Omega }/2)^2C_u(t)=0$$ (88) where $$\mathrm{\Omega }=\sqrt{2A_0\overline{A}(\omega _m)\delta \omega _m}.$$ (89) Hence we are left, in the approximation made, with a damped-oscillator equation of motion for the upper-state probability amplitude, where $`\overline{A}(\omega _m)`$ and $`\delta \omega _m`$, respectively, are given by Eqs. (79) and (80) \[or Eqs. (81) and (82)\]. Obviously, when $`\omega _A`$ $`=`$ $`\omega _m`$ and $$\mathrm{\Omega }\delta \omega _m$$ (90) (i.e., strong coupling), then damped Rabi oscillations are observed: $$\left|C_u(t)\right|^2=e^{\delta \omega _mt}\mathrm{cos}^2(\mathrm{\Omega }t/2).$$ (91) Note that in the opposite case where $`\mathrm{\Omega }`$ $``$ $`\delta \omega _m`$ the solution of Eq. (88) is $`|C_u(t)|^2`$ $`=`$ $`e^{At}`$ with $`A`$ from Eq. (74) for $`\omega _A`$ $`=`$ $`\omega _m`$, which is in agreement with Eq. (44). Equations (89) and (90) together with Eqs. (79) and (80) \[or Eqs. (81) and (82)\] provide us with an easy rule of thumb for deciding whether the strong-coupling regime is realized or not. In order to obtain the exact solution to the problem, we have also solved the basic integral equation (37) \[together with the kernel function (49)\] numerically. Typical examples of the temporal evolution of the occupation probability of the upper atomic state are shown in Fig. 5 for the case where the atomic transition is tuned to the resonance line in the center of the band gap. The figure reveals that with increasing value of the intrinsic damping constant $`\gamma `$ of the wall material the Rabi oscillations become less pronounced, in agreement with Eqs. (79), (80), (89), and the inequality (90). From Eqs. (79) and (80) it is seen that (in the approximation made there) the product $`\overline{A}(\omega _m)\delta \omega _m`$ $``$ $`c/R_2`$ does not vary with $`\gamma `$, whereas $`\delta \omega _m`$ linearly increases with $`\gamma `$. According to Eq. (89), $`\mathrm{\Omega }`$ $``$ $`\sqrt{\gamma }`$ and thus $`\mathrm{\Omega }/\delta \omega _m`$ $``$ $`1/\sqrt{\gamma }`$, i.e., the condition of strong coupling (90) becomes violated with increasing $`\gamma `$. Physically, increasing $`\gamma `$ means increasing probability of (irreversible) absorption of the emitted photon by the wall material and therefore reduced probability of (reversible) atom-field energy exchange. It should be mentioned that with increasing $`\gamma `$ the (very small) probability that the photon (irreversibly) leaves the cavity also increases. Note that the variation of $`\gamma `$ in Fig. 5 leaves the imaginary part of the refractive index nearly unchanged, $`n_\mathrm{I}`$ $``$ $`1.2`$, while the (small) real part $`n_\mathrm{R}`$ slightly increases with $`\gamma `$ (see Fig. 2). The examples of the temporal evolution of the occupation probability of the upper atomic state shown in Fig. 6 refer to the case where the atomic transition is tuned to a resonance line closest to $`\omega _\mathrm{T}`$ below the band gap. According to Eqs. (81), (82), (89), and the condition (90), the resonance frequencies close to $`\omega _\mathrm{T}`$ are most favorable for realizing the strong-coupling regime in the range of normal dispersion, because of the rising real part of the refractive index \[see Fig. 2(a)\]. As expected, the strength of the the Rabi oscillations now varies with the plasma frequency $`\omega _\mathrm{P}`$ such that they are less pronounced for small values of $`\omega _\mathrm{P}`$. Obviously, decreasing $`\omega _\mathrm{P}`$ means increasing input-output coupling, i.e., increasing probability that the emitted photon leaves the cavity instead of back-acting upon the atom. Finally, Fig. 6 presents a comparison between the exact solution, and the approximate analytical solution \[solution of Eq. (88)\]. To compensate for the short-time inaccuracy of the analytical solution, it is somewhat shifted forward in time. The agreement between the exact solution and the (shifted) analytical solution is quite good. Obviously, Eq. (88) predicts an initial decay somewhat faster than the exact one. In fact, the spontaneous decay is accelerated only gradually under the back-action of the radiated field being multiply reflected at the boundaries , and the single-resonance coupling is established only after a certain interval of time. The sharper is the cavity resonance, the shorter is this interval, which is fully confirmed by the figure. ## IV Summary and Conclusions We have developed a formalism for studying spontaneous decay of an excited atom in the presence of arbitrary dispersing and absorbing dielectric bodies. The formalism is based on a source-quantity representation of the electromagnetic field in terms of the Green tensor of the classical problem and appropriately chosen bosonic quantum fields. It replaces the standard concept of mode decomposition which fails for complex permittivity. All relevant information about the bodies such as form and intrinsic dispersion and absorption properties are contained in the Green tensor. It is worth noting that the Green tensor has been available for a large variety of configurations such as planarly, spherically, and cylindrically multilayered media . We have applied the theory to the spontaneous decay of a two-level atom placed at the center of a three-layered spherical microcavity, modeling the wall by a Lorentz dielectric. The formalism has enabled us to study both the range of normal dispersion and the anomalous-dispersion range within the band gap in a unified way. Whereas in the range of normal dispersion the cavity input–output coupling dominates the strength of the atom-field interaction, the dominating effect within the band gap is the photon absorption by the wall material. In the study of the spherical-cavity problem, we have assumed that the atom is placed at the center of the cavity, which has drastically reduced the mathematical effort, because only the spherical Bessel function of order $`n`$ $`=`$ $`1`$ contributes to the Green tensor. The price we paid is that the interaction of the atom with cavity excitations of large $`n`$, which correspond to high-$`Q`$ whispering gallery modes and concentrate near the surface by repeated internal reflections has not been included in the analysis. Since the complete Green tensor is known, there is of course no obstacle to perform the calculations for an arbitrary position of the atom. In particular, when the atom is near the surface then nonradiative energy transfer from the atom to the absorbing medium can substantially contribute to the process of spontaneous decay. Further investigations are also necessary in order to give a more detailed analysis of the evolution of the emitted radiation, to answer the question of the ratio of photon emission to nonradiative decay, and to extend the theory to multilevel atom-field interactions. ###### Acknowledgements. We thank Stefan Scheel for helpful discussions. H.T.D. is grateful to the Alexander von Humboldt-Stiftung for a research fellowship. This work was supported by the Deutsche Forschungsgemeinschaft. ## A The Hamiltonian in the dipole and rotating wave approximations The Hamiltonian (21) can be rewritten as $`\widehat{H}=\widehat{H}_\mathrm{F}+\widehat{H}_\mathrm{A}+\widehat{H}_{\mathrm{AF}},`$ (A1) $`\widehat{H}_\mathrm{F}={\displaystyle d^3𝐫_0^{\mathrm{}}𝑑\omega \mathrm{}\omega \widehat{𝐟}^{}(𝐫,\omega )\widehat{𝐟}(𝐫,\omega )},`$ (A2) $`\widehat{H}_\mathrm{A}={\displaystyle \underset{\alpha }{}}{\displaystyle \frac{\widehat{𝐩}_\alpha ^2}{2m_\alpha }}+{\displaystyle \frac{1}{2}}{\displaystyle d^3𝐫\widehat{\rho }_\mathrm{A}(𝐫)\widehat{\phi }_\mathrm{A}(𝐫)},`$ (A3) $`\widehat{H}_{\mathrm{AF}}={\displaystyle \underset{\alpha }{}}{\displaystyle \frac{q_\alpha }{m_\alpha }}\widehat{𝐩}_\alpha \widehat{𝐀}(𝐫_\alpha )+{\displaystyle d^3𝐫\widehat{\rho }_\mathrm{A}(𝐫)\widehat{\phi }(𝐫)},`$ (A4) where we have ignored the small $`\widehat{𝐀}^2`$ term. Note that in the Coulomb gauge $`[\widehat{𝐩}_\alpha ,\widehat{𝐀}]=0`$. For a neutral atom with the nucleus being positioned at $`𝐫_\mathrm{A}`$, the atomic dipole operator reads $$\widehat{𝝁}_\mathrm{A}\underset{\alpha }{}q_\alpha (\widehat{𝐫}_\alpha 𝐫_\mathrm{A})=\underset{\alpha }{}q_\alpha \widehat{𝐫}_\alpha ,$$ (A5) and the first term in the interaction part of the Hamiltonian $`\widehat{H}_{\mathrm{AF}}`$ takes the form $`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{q_\alpha }{m_\alpha }}\widehat{𝐩}_\alpha \widehat{𝐀}(𝐫_\alpha )`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{q_\alpha }{i\mathrm{}}}[\widehat{𝐫}_\alpha ,\widehat{H}_\mathrm{A}]\widehat{𝐀}(𝐫_\mathrm{A})`$ (A7) $`={\displaystyle \frac{1}{i\mathrm{}}}[\widehat{𝝁}_\mathrm{A},\widehat{H}_\mathrm{A}]\widehat{𝐀}(𝐫_\mathrm{A}),`$ where the dipole approximation has been employed to replace $`\widehat{𝐀}(𝐫_\alpha )`$ $``$ $`\widehat{𝐀}(𝐫_\mathrm{A})`$ and $`\widehat{𝐩}_\alpha `$ $`=`$ $`(m_\alpha /(i\mathrm{}))[\widehat{𝐫}_\alpha ,\widehat{H}_\mathrm{A}]`$ has been used. Now we restrict ourselves to a two-state model of an atom with upper state $`|u`$ and lower state $`|l`$. These are eigenstates of the unperturbed part of the Hamiltonian $`\widehat{H}_\mathrm{A}`$ with the eigenvalues $`\mathrm{}\omega _u`$ and $`\mathrm{}\omega _l`$, respectively. Then one can write $`\widehat{H}_\mathrm{A}=\mathrm{}\omega _u|uu|+\mathrm{}\omega _l|ll|,`$ (A8) $`|uu|+|ll|=\widehat{I}.`$ (A9) In this atomic state space, the dipole operator $`\widehat{𝝁}_\mathrm{A}`$ has the matrix elements $`u|\widehat{𝝁}_\mathrm{A}|l=l|\widehat{𝝁}_\mathrm{A}|u𝝁`$ and $`u|\widehat{𝝁}_\mathrm{A}|u=l|\widehat{𝝁}_\mathrm{A}|l=0`$. Using them and Eqs. (17), (A7) – (A9), we arrive at $`{\displaystyle \underset{\alpha }{}}{\displaystyle \frac{q_\alpha }{m_\alpha }}\widehat{𝐩}_\alpha \widehat{𝐀}(𝐫_\alpha )`$ (A12) $`(\widehat{\sigma }\widehat{\sigma }^{})[{\displaystyle _0^{\mathrm{}}}d\omega {\displaystyle \frac{\omega _\mathrm{A}}{\omega }}\underset{¯}{\overset{^}{𝐄}}^{}(𝐫_\mathrm{A},\omega )\mathrm{H}.\mathrm{c}.]𝝁`$ $`[\widehat{\sigma }^{}\widehat{𝐄}^{(+)}(𝐫_\mathrm{A})𝝁+\mathrm{H}.\mathrm{c}.],`$ where $`\widehat{\sigma }=|lu|,\widehat{\sigma }^{}=|ul|`$, $`\omega _\mathrm{A}=\omega _u\omega _l`$, and $`\omega =\omega _\mathrm{A}`$ is set in the integral, because of the rotating wave approximation. In order to deal with the second term in $`\widehat{H}_{\mathrm{AF}}`$, we expand $`\widehat{\rho }_\mathrm{A}(𝐫)`$ in a multi-polar form and retain only the first non-vanishing term $`\widehat{\rho }_\mathrm{A}(𝐫){\displaystyle \underset{\alpha }{}}q_\alpha \delta (𝐫𝐫_\mathrm{A})`$ (A15) $`\mathbf{}\left[\delta (𝐫𝐫_\mathrm{A}){\displaystyle \underset{\alpha }{}}q_\alpha (\widehat{𝐫}_\alpha 𝐫_\mathrm{A})\right]`$ $`=\mathbf{}\delta (𝐫𝐫_\mathrm{A})\widehat{𝝁}_\mathrm{A}.`$ Then we have $`{\displaystyle d^3𝐫\widehat{\rho }_\mathrm{A}(𝐫)\widehat{\phi }(𝐫)}{\displaystyle d^3𝐫\left\{\mathbf{}[\delta (𝐫𝐫_\mathrm{A})\widehat{𝝁}_\mathrm{A}]\right\}\widehat{\phi }(𝐫)}`$ (A19) $`={\displaystyle d^3𝐫[\delta (𝐫𝐫_\mathrm{A})\widehat{𝝁}_\mathrm{A}]\mathbf{}\widehat{\phi }(𝐫)}`$ $`=\widehat{𝝁}_\mathrm{A}\widehat{𝐄}^{}(𝐫_\mathrm{A})`$ $`[\widehat{\sigma }^{}\widehat{𝐄}^{(+)}(𝐫_\mathrm{A})𝝁+\mathrm{H}.\mathrm{c}.],`$ where integration by parts and Eq. (16) have been employed for deriving the second and the third equation, respectively, and the rotating wave approximation has been used in deriving the forth equation. Combining (A4), (A12), and (A19) gives $$\widehat{H}_{\mathrm{AF}}[\widehat{\sigma }^{}\widehat{𝐄}^{(+)}(𝐫_\mathrm{A})𝝁+\mathrm{H}.\mathrm{c}.].$$ (A20) Equations (A1), (A2), (A20), and a subtraction of $`(\mathrm{}/2)(\omega _u+\omega _l)\widehat{I}`$ from $`\widehat{H}_\mathrm{A}`$, Eq. (A8), lead to the Hamiltonian (25). ## B The Green tensor of the spherical cavity Following, we write the Green tensor of the cavity in Fig. 1 in the form $$𝑮(𝐫,𝐫^{},\omega )=𝑮^\mathrm{V}(𝐫,𝐫^{},\omega )\delta _{fs}+𝑮^{(fs)}(𝐫,𝐫^{},\omega ),$$ (B1) where $`𝑮^\mathrm{V}(𝐫,𝐫^{},\omega )`$ represents the contribution of the direct waves from the radiation sources in an unbounded medium, which is vacuum in our case, $`f`$ and $`s`$ denote the layers where the field point and source point locate, $`\delta _{fs}`$ is the usual Kronecker symbol, and the scattering Green tensor $`𝑮^{(fs)}(𝐫,𝐫^{},\omega )`$ describes the contribution of the multiple reflection ($`f=s`$) and transmission ($`fs`$) waves. In particular, $`𝑮^{(13)}(𝐫,𝐫^{},\omega )`$ and $`𝑮^{(33)}(𝐫,𝐫^{},\omega )`$ read as $`𝑮^{(13)}(𝐫,𝐫^{},\omega )={\displaystyle \frac{ik_3}{4\pi }}{\displaystyle \underset{e,o}{}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=0}{\overset{n}{}}}`$ (B5) $`\times \{{\displaystyle \frac{2n+1}{n(n+1)}}{\displaystyle \frac{(nm)!}{(n+m)!}}(2\delta _{0m})`$ $`\times [A_M^{13}(\omega )𝐌_{\genfrac{}{}{0pt}{}{e}{o}nm}^{(1)}(𝐫,k_1)𝐌_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫^{},k_3)`$ $`+A_N^{13}(\omega )𝐍_{\genfrac{}{}{0pt}{}{e}{o}nm}^{(1)}(𝐫,k_1)𝐍_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫^{},k_3)]\},`$ $`𝑮^{(33)}(𝐫,𝐫^{},\omega )={\displaystyle \frac{ik_3}{4\pi }}{\displaystyle \underset{e,o}{}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m=0}{\overset{n}{}}}`$ (B9) $`\times \{{\displaystyle \frac{2n+1}{n(n+1)}}{\displaystyle \frac{(nm)!}{(n+m)!}}(2\delta _{0m})`$ $`\times [C_M^{33}(\omega )𝐌_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫,k_1)𝐌_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫^{},k_3)`$ $`+C_N^{33}(\omega )𝐍_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫,k_1)𝐍_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫^{},k_3)]\},`$ where $$k_1=k_3=\frac{\omega }{c},k_2=\sqrt{ϵ(\omega )}\frac{\omega }{c}$$ and $`𝐌`$ and $`𝐍`$ represent TM- and TE-waves, respectively, $`𝐌_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫,k)`$ $`={\displaystyle \frac{m}{\mathrm{sin}\theta }}j_n(kr)P_n^m(\mathrm{cos}\theta )\left({\displaystyle \genfrac{}{}{0pt}{}{\mathrm{sin}}{\mathrm{cos}}}\right)(m\varphi )𝐞_\theta `$ (B11) $`j_n(kr){\displaystyle \frac{dP_n^m(\mathrm{cos}\theta )}{d\theta }}\left({\displaystyle \genfrac{}{}{0pt}{}{\mathrm{cos}}{\mathrm{sin}}}\right)(m\varphi )𝐞_\varphi ,`$ $`𝐍_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫,k)`$ $`={\displaystyle \frac{n(n+1)}{kr}}j_n(kr)P_n^m(\mathrm{cos}\theta )\left({\displaystyle \genfrac{}{}{0pt}{}{\mathrm{cos}}{\mathrm{sin}}}\right)(m\varphi )𝐞_r`$ (B14) $`+{\displaystyle \frac{1}{kr}}{\displaystyle \frac{d[rj_n(kr)]}{dr}}[{\displaystyle \frac{dP_n^m(\mathrm{cos}\theta )}{d\theta }}\left({\displaystyle \genfrac{}{}{0pt}{}{\mathrm{cos}}{\mathrm{sin}}}\right)(m\varphi )𝐞_\theta `$ $`{\displaystyle \frac{m}{\mathrm{sin}\theta }}P_n^m(\mathrm{cos}\theta )\left({\displaystyle \genfrac{}{}{0pt}{}{\mathrm{sin}}{\mathrm{cos}}}\right)(m\varphi )𝐞_\varphi ],`$ with $`j_n(x)`$ being the spherical Bessel function of the first kind and $`P_n^m(x)`$ being the associated Legendre function. The superscript $`(1)`$ in Eq. (B5) indicates that in Eqs. (B11) and (B14) the spherical Bessel function $`j_n(x)`$ has to be replaced by the first-type spherical Hankel function $`h_n^{(1)}(x)`$. The coefficients $`A_{M,N}^{13}`$ and $`C_{M,N}^{33}`$ in Eqs. (B5) and (B9) are defined by $`A_{M,N}^{13}(\omega )`$ $`={\displaystyle \frac{T_{F1}^{M,N}T_{F2}^{M,N}T_{P1}^{M,N}}{T_{P1}^{M,N}+T_{F1}^{M,N}R_{P1}^{M,N}R_{F2}^{M,N}}},`$ (B15) $`C_{M,N}^{33}(\omega )`$ $`={\displaystyle \frac{A_{M,N}^{13}}{T_{P2}^{M,N}}}\left[{\displaystyle \frac{R_{P2}^{M,N}}{T_{F1}^{M,N}}}+{\displaystyle \frac{R_{P1}^{M,N}}{T_{P1}^{M,N}}}\right],`$ (B16) where $`R_{Pf}^M`$ $`={\displaystyle \frac{k_{f+1}H_{(f+1)f}^{}H_{ff}k_fH_{ff}^{}H_{(f+1)f}}{k_{f+1}J_{ff}H_{(f+1)f}^{}k_fJ_{ff}^{}H_{(f+1)f}}},`$ (B17) $`R_{Ff}^M`$ $`={\displaystyle \frac{k_{f+1}J_{(f+1)f}^{}J_{ff}k_fJ_{ff}^{}J_{(f+1)f}}{k_{f+1}J_{(f+1)f}^{}H_{ff}k_fJ_{(f+1)f}H_{ff}^{}}},`$ (B18) $`R_{Pf}^N`$ $`={\displaystyle \frac{k_{f+1}H_{(f+1)f}H_{ff}^{}k_fH_{ff}H_{(f+1)f}^{}}{k_{f+1}J_{ff}^{}H_{(f+1)f}k_fJ_{ff}H_{(f+1)f}^{}}},`$ (B19) $`R_{Ff}^N`$ $`={\displaystyle \frac{k_{f+1}J_{(f+1)f}J_{ff}^{}k_fJ_{ff}J_{(f+1)f}^{}}{k_{f+1}J_{(f+1)f}H_{ff}^{}k_fJ_{(f+1)f}^{}H_{ff}}},`$ (B20) $`T_{Pf}^M`$ $`={\displaystyle \frac{k_{f+1}\left(J_{(f+1)f}H_{(f+1)f}^{}J_{(f+1)f}^{}H_{(f+1)f}\right)}{k_{f+1}J_{ff}H_{(f+1)f}^{}k_fJ_{ff}^{}H_{(f+1)f}}},`$ (B21) $`T_{Ff}^M`$ $`={\displaystyle \frac{k_{f+1}\left(J_{(f+1)f}^{}H_{(f+1)f}J_{(f+1)f}H_{(f+1)f}^{}\right)}{k_{f+1}J_{(f+1)f}^{}H_{ff}k_fJ_{(f+1)f}H_{ff}^{}}},`$ (B22) $`T_{Pf}^N`$ $`={\displaystyle \frac{k_{f+1}\left(J_{(f+1)f}^{}H_{(f+1)f}J_{(f+1)f}H_{(f+1)f}^{}\right)}{k_{f+1}J_{ff}^{}H_{(f+1)f}k_fJ_{ff}H_{(f+1)f}^{}}},`$ (B23) $`T_{Ff}^N`$ $`={\displaystyle \frac{k_{f+1}\left(J_{(f+1)f}H_{(f+1)f}^{}J_{(f+1)f}^{}H_{(f+1)f}\right)}{k_{f+1}J_{(f+1)f}H_{ff}^{}k_fJ_{(f+1)f}^{}H_{ff}}}`$ (B24) with $`J_{il}`$ $`=j_n(k_iR_l),`$ (B25) $`H_{il}`$ $`=h_n^{(1)}(k_iR_l),`$ (B26) $`J_{il}^{}`$ $`={\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d[\rho j_n(\rho )]}{d\rho }}|_{\rho =k_iR_l},`$ (B27) $`H_{il}^{}`$ $`={\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d[\rho h_n^{(1)}(\rho )]}{d\rho }}|_{\rho =k_iR_l}.`$ (B28) Note that $`A_{M,N}^{13}`$ and $`C_{M,N}^{33}`$ are functions of $`n`$ but not of $`m`$. When the atom is positioned at the cavity center, we have $`𝐌_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫,k)|_{kr0}(kr)^n,`$ (B29) $`𝐍_{\genfrac{}{}{0pt}{}{e}{o}nm}(𝐫,k)|_{kr0}(kr)^{n1}.`$ (B30) In this case, only TM-waves with $`n`$ $`=`$ $`1`$ contribute and the Eq. (B9) simplifies to $$𝑮^\mathrm{R}𝑮^{(33)}(𝐫,𝐫^{},\omega )|_{r=r^{}0}=\frac{i\omega }{6\pi c}C_N^{33}(\omega )𝑰.$$ (B31) Similarly, Eq. (B5) reduces to $`G_{rz}^{(13)}(𝐫,𝐫^{},\omega )|_{r^{}0}={\displaystyle \frac{i\mathrm{cos}\theta }{2\pi r}}h_1^{(1)}(k_3r)A_N^{13}(\omega ),`$ (B35) $`G_{\theta z}^{(13)}(𝐫,𝐫^{},\omega )|_{r^{}0}={\displaystyle \frac{i\mathrm{sin}\theta }{4\pi r}}{\displaystyle \frac{d[rh_1^{(1)}(k_3r)]}{dr}}A_N^{13}(\omega ),`$ $`G_{\varphi z}^{(13)}(𝐫,𝐫^{},\omega )|_{r^{}0}=0.`$ ## C Derivation of Eq. (84) For an atom at the cavity center and $`z`$-oriented dipole, from Eqs. (61) and (B35) – (B35) we derive $$F_r(𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})=𝒪(\rho ^2),$$ (C1) $`F_\theta (𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})={\displaystyle \frac{k_\mathrm{A}^2\mu \mathrm{sin}\theta }{4\pi ^2ϵ_0\rho }}`$ (C3) $`\times {\displaystyle _0^{\mathrm{}}}d\omega \mathrm{Im}\left[A_N^{13}(\omega )e^{i\omega \rho /c}\right]\zeta (\omega _\mathrm{A}\omega )+𝒪(\rho ^2),`$ $$F_\varphi (𝐫,𝐫_\mathrm{A},\omega _\mathrm{A})=0.$$ (C4) Recalling the relation $`\zeta (x)`$ $`=`$ $`i/(x+i0)`$, we perform the $`\omega `$-integration in Eq. (C3) to obtain $`{\displaystyle _0^{\mathrm{}}}𝑑\omega \mathrm{Im}\left[A_N^{13}(\omega )e^{i\omega \rho /c}\right]\zeta (\omega _\mathrm{A}\omega )`$ (C7) $`={\displaystyle \frac{1}{2}}{\displaystyle _0^{\mathrm{}}}𝑑\omega {\displaystyle \frac{\{A_N^{13}(\omega )e^{i\omega \rho /c}\mathrm{c}.\mathrm{c}.\}}{\omega (\omega _\mathrm{A}+i0)}}`$ $`i\pi A_N^{13}(\omega _\mathrm{A})e^{i\omega _\mathrm{A}\rho /c},`$ where we have (approximately) replaced $`A_N^{13}(\omega )`$ by $`A_N^{13}(\omega _\mathrm{A})`$, extended the lower limit of the integral to $`\mathrm{}`$, and applied contour-integration techniques. Combining Eqs. (59), (61), (83), (C1), (C4), and (C7), it is not difficult to prove that $$W\mathrm{}\omega _\mathrm{A}\frac{|A_N^{13}(\omega _\mathrm{A})|^2}{1+\mathrm{Re}C_N^{33}(\omega _\mathrm{A})}.$$ (C8) Taking into account that the free-space value $`W_0`$ $`=`$ $`\mathrm{}\omega _\mathrm{A}`$ is obtained by setting $`A_N^{13}`$ $`=`$ $`1`$ and $`C_N^{33}`$ $`=`$ $`0`$, Eq. (C8) just yields Eq. (84).
warning/0003/hep-ph0003306.html
ar5iv
text
# Mini-review on Collider Signatures for Extra Dimensions ## 1 Introduction Recent advances in string theory have revolutionized particle phenomenology. Namely, the previously unreachable Planck, string, and grand unification scales ($`M_{\mathrm{Pl}}`$, $`M_{\mathrm{st}}`$, and $`M_{\mathrm{GUT}}`$, respectively) can be brought down to a TeV range through the existence of extra dimensions. There have been a number of ideas that can bring either the Planck, GUT, or string scale down to TeV region. One expects the low energy phenomenology of these new ideas or models can be tested at current and future collider experiments. In this talk, we review collider signatures for the following two models of extra dimensions: (i) the one proposed by Arkani-Hamed, Dimopoulos, and Dvali , in which the standard model (SM) particles live on a 3-brane while the gravity is free to propagate in extra dimensions of very large size ($`\stackrel{<}{}1`$ mm). This model was motivated by the fact that the effective Planck scale is brought down to TeV to solve the hierarchy problem. Collider signatures for this model can be divided into two types: (a) emission of real gravitons into extra dimensions and thus gives rise to missing energy signals, and (b) exchanges of virtual gravitons that frequently lead to enhancement of production of SM particles. (ii) The model proposed by Dienes et al. and Pomarol et al. that the SM gauge bosons are allowed to propagate in extra dimensions (whereas the gravity effect here is negligible.) This model has the merit of unifying the gauge couplings at a scale much lower than the usual GUT scale. Collider signatures for this model are due to the existence of the Kaluza-Klein (KK) states of $`\gamma `$, $`W`$, $`Z`$, and $`g`$ bosons. ## 2 Model of Arkani-Hamed, Dimopoulos, and Dvali This model was first proposed to solve the hierarchy problem by requiring the compactified dimensions to be of very large size, $`\stackrel{<}{}1`$ mm. While the SM particles live on a D3-brane, the gravity is free to propagate in extra dimensions. Using Gauss law, the effective Planck scale $`M_S`$ is related to the four-dimensional Planck scale $`M_{\mathrm{Pl}}`$ ($`10^{19}`$ GeV) by $$M_{\mathrm{Pl}}^2M_S^{n+2}R^n,$$ where $`n`$ is the number of extra (compactified) dimensions and $`R`$ is the size of the compactified dimensions. Assuming that the effective Planck scale $`M_S`$ is in the TeV range, it gives a very large $`R`$ of the size of a solar system for $`n=1`$, which is obviously ruled out by experiments. However, for all $`n2`$ the expected $`R`$ is less than $`1`$ mm, and therefore do not contradict existing gravitational experiments. With the SM particles residing on the brane and the graviton freely propagating in extra dimensions, in the 4D-point of view a graviton in extra dimensions is equivalent to a tower of infinite number of KK states with masses $`M_k=2\pi k/R(k=0,1,2,\mathrm{},\mathrm{})`$. The couplings of SM particles to each of these KK states is still of order $`1/M_{\mathrm{Pl}}`$, but the overall coupling is, however, obtained by summing over all the KK states, and therefore scales as $`1/M_S`$. Since the $`M_S`$ is in the TeV range, the effective gravitational interaction is as strong as the electroweak interaction and thus gives rise to many testable consequences in both accelerator and non-accelerator experiments. A large number of phenomenological studies in this area have appeared. The collider signatures can be divided into two categories: (i) real emission of gravitons into extra dimensions, giving rise to missing energy signal, and (ii) virtual exchange of gravitons in addition to exchanges of SM gauge bosons, giving rise to enhancement or deviations from the SM predictions. We summarize these signals in the following. Note that a stringent constraint comes from astrophysical (SN1987A) and cosmological sources and the lower bound on the effective Planck scale $`M_S`$ is 30–100 TeV for $`n=2`$. ### 2.1 Real Emission of Gravitons Since gravitons interact weakly with detectors, they will escape detection and causing missing energies. Thus, the logical signal to search for would be the associated production of gravitons with other SM particles. At $`e^+e^{}`$ colliders, the best signals would be the associated production of graviton with a $`Z`$ boson, a photon, or a fermion pair. The production of graviton and photon at LEPII has been studied . The striking signature would be a single photon with missing energy while the irreducible background comes from the process $`e^+e^{}\gamma \nu _{\mathrm{}}\overline{\nu }_{\mathrm{}},(\mathrm{}=e,\mu ,\tau )`$. The cross section for the signal is given by $`{\displaystyle \frac{d\sigma }{d\mathrm{cos}\theta }}`$ $`=`$ $`{\displaystyle \frac{\pi \alpha G_N}{4\left(1\frac{m^2}{s}\right)}}[(1+\mathrm{cos}^2\theta )(1+({\displaystyle \frac{m^2}{s}})^4)`$ (1) $`+`$ $`\left({\displaystyle \frac{13\mathrm{cos}^2\theta +4\mathrm{cos}^4\theta }{1\mathrm{cos}^2\theta }}\right){\displaystyle \frac{m^2}{s}}(1+({\displaystyle \frac{m^2}{s}})^2)+6\mathrm{cos}^2\theta ({\displaystyle \frac{m^2}{s}})^2].`$ The signal cross section increases with the energy of collision while the background is gradually decreasing after the $`Z`$-peak. At LEPII, if the effective Planck scale $`M_S`$ is low enough deviations from the SM prediction should be seen. In a recent search by L3 , the limit $`M_S\stackrel{>}{}1`$ TeV for $`n=2`$. The production of graviton with a $`Z`$ boson at LEPII gives a signature of a $`Z`$ boson, which decays into a $`q\overline{q}`$ or $`\mathrm{}\overline{\mathrm{}}`$ pair, plus missing energy. This is an interesting process because LEPII already searched for the invisibly decaying Higgs boson in $`ZH`$ production, which has the same signature as the $`ZG`$ production. The formulas for the signal cross section can be found in Ref. . The irreducible background is $`e^+e^{}Z\nu \overline{\nu }`$. Based on the existing data on the search for invisibly decaying Higgs, the lower limit on $`M_S`$ is around 515 GeV. A refined search by L3 gives a limit $`M_S>600`$ GeV. The associated production of graviton with a $`f\overline{f}`$ pair was studied at $`Z^0`$ because of a large number of hadronic $`Z`$ decays. The signature would be a fermion pair with missing energy. The background comes from $`Zf\overline{f}\nu \overline{\nu }`$, which has a small BR of $`210^7`$. The present data agrees with the SM prediction and is able to constrain $`M_S`$ to be at least 0.4 TeV for $`n=2`$ . Another exciting opportunity is the monojet plus missing energy production at hadronic colliders. ### 2.2 Virtual Exchanges of gravitons There are numerous studies in collider signatures associated with virtual exchanges of KK-gravitons, including diphoton, diboson, and fermion-pair production. In the following, we highlight some of these studies. One of the most prominent channels is photon-photon scattering. In the SM, photon-photon scattering only takes place via box diagrams of $`W`$ bosons and quarks so that it is loop-suppressed. On the other hand, in the ADD model photons can scatter via exchanges of gravitons in $`s,t,u`$-channels. The scattering amplitude-squared is symmetric in $`s,t,u`$ and given by $`\overline{}||^2=\frac{\kappa ^4}{8}|D(s)|^2(s^4+t^4+u^4)`$. The differential cross section is given by $$\frac{d\sigma (\gamma \gamma \gamma \gamma )}{d|\mathrm{cos}\theta |}=\frac{\pi s^3}{M_S^8}^2[1+\frac{1}{8}(1+6\mathrm{cos}^2\theta +\mathrm{cos}^4\theta )],$$ where $`=2/(n2)`$ for $`n>2`$. The signal cross section easily surpasses the SM background at around $`\sqrt{s}=0.5`$ TeV for $`M_S=4`$ TeV . The polarized scattering has also been studied in Ref. . Another interesting channel is neutrino-photon scattering . Diphoton production and also $`WW,ZZ`$ production at $`e^+e^{}`$ colliders have been studied. The effect of TeV scale gravity on the angular distribution of diphoton production is $$\frac{d\sigma (e^+e^{}\gamma \gamma )}{dz}=\frac{2\pi }{s}(\alpha \sqrt{\frac{1+z^2}{1z^2}}+\frac{s^2}{8}\frac{}{M_S^4}\sqrt{1z^4})^2$$ (2) where $`z=|\mathrm{cos}\theta _\gamma |`$. This effect is very similar to the conventional deviation from QED, which is often parametrized by a $`\mathrm{\Lambda }`$ as $$\frac{d\sigma }{dz}=\frac{2\pi \alpha ^2}{s}\frac{1+z^2}{1z^2}\left(1\pm \frac{s^2}{2\mathrm{\Lambda }_\pm ^4}(1z^2)\right).$$ We can immediately equate the above two expressions and arrive at $`\frac{M_S^4}{}=\frac{\mathrm{\Lambda }_+^4}{2\alpha }`$. The present limits obtained by LEP Collaborations on $`\mathrm{\Lambda }_{\mathrm{QED}}262345`$ GeV, which is equivalent to $`M_S0.71`$ TeV for $`n=4`$ . Agashe and Deshpande also studied $`e^+e^{}\gamma \gamma ,W^+W^{},ZZ`$ production and compared their sensitivities to TeV scale gravity. Interestingly, $`ZZ`$ production offers the highest fractional change of cross section among $`\gamma \gamma ,WW,ZZ`$ due to gravity effects. However, the $`ZZ`$ production rate is smaller than the other two. Overall, their sensitivities are similar. A recent experimental search performed by L3 found that the sensitivities of $`\gamma \gamma ,WW,ZZ`$ are very similar and the combined limit is $`M_S\stackrel{>}{}0.8`$ TeV. Diphoton production is one of the best probes of TeV scale gravity at hadron colliders. The angular distributions of the subprocesses are given by $`{\displaystyle \frac{d\sigma (q\overline{q}\gamma \gamma )}{d\mathrm{cos}\theta ^{}}}`$ $`=`$ $`{\displaystyle \frac{1}{48\pi \widehat{s}}}[e^2Q_q^2\sqrt{{\displaystyle \frac{1+\mathrm{cos}^2\theta ^{}}{1\mathrm{cos}^2\theta ^{}}}}+{\displaystyle \frac{\pi \widehat{s}^2}{2}}{\displaystyle \frac{}{M_S^4}}\sqrt{1\mathrm{cos}^4\theta ^{}}]^2,`$ (3) $`{\displaystyle \frac{d\sigma (gg\gamma \gamma )}{d\mathrm{cos}\theta ^{}}}`$ $`=`$ $`{\displaystyle \frac{\pi \widehat{s}^3}{512}}\left({\displaystyle \frac{}{M_S^4}}\right)^2(1+6\mathrm{cos}^2\theta ^{}+\mathrm{cos}^4\theta ^{}),`$ (4) where $`\mathrm{cos}\theta ^{}`$ is the scattering angle of the photon in the center-of-mass frame of the incoming partons, and here $`\mathrm{cos}\theta ^{}`$ is from $`1`$ to 1. Based on the existing diphoton data from CDF that in $`M_{\gamma \gamma }>150`$ GeV 5 events are observed where $`4.5\pm 0.6`$ are expected with a luminosity of 100 pb<sup>-1</sup>, a limit of $`M_S>0.9`$ TeV for $`n=4`$ was obtained. The upcoming CDF and D0 searches will easily overshadow this limit. The general vector-boson scattering $`VVVV`$, where $`V=\gamma ,W,Z`$ was studied by Atwood et al. . The conclusion is that the effect of TeV scale gravity shows up at large invariant mass region. Extra dimensions also affect fermion-pair production at $`e^+e^{}`$ colliders and the corresponding crossing channels, such as Drell-Yan production at the Tevatron and neutral-current (NC) deep-inelastic scattering (DIS) at HERA. While they were individually studied in a number of publications , a comphrensive analysis of all these data sets was performed in Ref. . The combined limit on $`M_S`$ is $`M_S>0.94`$ TeV for $`n=4`$. Bourilkov , on the other hand, used the combined data of Bhabha scattering of the four LEP Collaborations and was able to obtain a limit of $`M_S>1.4`$ TeV. There are also a combined search in fermion-pair production, diphoton, $`WW`$ and $`ZZ`$ pair production by L3 that a limit of $`M_S\stackrel{>}{}1`$ TeV is established. Dijet and top-pair production at the Tevatron or other colliders should also be useful in obtaining information on $`M_S`$, but systematics will likely reduce the usefulness of these channels. Effects on precision variables and effects on patterns of fermion or neutrino masses have also been studied. Interactions with scalars or Higgs bosons have been investigated in Ref. . In the following, we describe a few studies that test sensitivity to $`M_S`$ in future experiments at hadronic and $`e^+e^{}`$ colliders. Cheung and Landsberg used double differential cross-sections, $`d^2\sigma /dMd\mathrm{cos}\theta ^{}`$, of diphoton and Drell-Yan production to constrain the effective Planck scale $`M_S`$ in Run I and Run II at the Tevatron and at the LHC. The advantage of using double differential distributions is that the differences in invariant mass and scattering angle between the SM and the gravity model can be contrasted simultaneously. Furthermore, for a $`22`$ process the invariant mass $`M`$ and the central scattering angle $`\mathrm{cos}\theta ^{}`$ already span the entire phase space. We, therefore, do not need further optimization of cuts or variables. The resulting sensitivities to $`M_S`$ are shown in Table 1(a). The sensitivity reach at linear $`e^+e^{}`$ colliders was studied in a number of work . Here I present a work that uses diphoton, Bhabha scattering, $`\mu ^+\mu ^{}`$, $`\tau ^+\tau ^{}`$, $`q\overline{q}`$ production and their angular distributions. The sensitivity limits on $`M_S`$ are obtained by a combined maximum likelihood approach by adding the likelihoods of all channels. The Bhabha scattering turns out to be the dominant channel. The angular distribution for diphoton production is given in Eq. (2) and for fermion-pair production it is given by $`{\displaystyle \frac{d\sigma (e^{}e^+f\overline{f})}{dz}}`$ $`=`$ $`{\displaystyle \frac{N_fs}{128\pi }}\{(1+z)^2(|M_{LL}(s)|^2+|M_{RR}(s)|^2)+(1z)^2`$ $`\times (|M_{RL}(s)|^2+|M_{LR}(s)|^2)+\pi ^2s^2(13z^2+4z^4)\eta ^2`$ $`\mathrm{\hspace{0.33em}8}\pi e^2Q_eQ_fz^3\eta +{\displaystyle \frac{8\pi e^2}{s_\theta ^2c_\theta ^2}}{\displaystyle \frac{s}{sM_Z^2}}(g_a^eg_a^f{\displaystyle \frac{13z^2}{2}}g_v^eg_v^fz^3)\eta \}`$ $`+{\displaystyle \frac{\delta _{ef}s}{128\pi }}\{(1+z)^2(|M_{LL}(t)|^2+|M_{RR}(t)|^2+2M_{LL}(s)M_{LL}(t)`$ $`+\mathrm{\hspace{0.33em}2}M_{RR}(s)M_{RR}(t))+4(|M_{LR}(t)|^2+|M_{RL}(t)|^2)+{\displaystyle \frac{\pi ^2s^2}{8}}`$ $`\times (121+228z+198z^2+84z^3+9z^4)\eta ^2{\displaystyle \frac{\pi s}{2}}\eta (M_{LL}(t)+M_{RR}(t)`$ $`+M_{LL}(s)+M_{RR}(s))(1+z)^2(7+z)+\pi s\eta (M_{LL}(t)+M_{RR}(t))`$ $`\times (1+z)^2(12z)2\pi s\eta (M_{LR}(t)+M_{RL}(t))(5+3z)\}`$ where $`\eta =/M_S^4`$ and the reduced amplitudes are given by $$M_{\alpha \beta }^{ef}(s)=\frac{e^2Q_eQ_f}{s}+\frac{e^2}{\mathrm{sin}^2\theta _\mathrm{w}\mathrm{cos}^2\theta _\mathrm{w}}\frac{g_\alpha ^eg_\beta ^f}{sM_Z^2},\alpha ,\beta =L,R.$$ The 95% C.L. sensitivity limits on $`M_S`$ for $`n=4`$ are shown in Table 1(b). The values for other $`n`$ can be easily obtained by scaling with $`=2/(n2)`$ for $`n>2`$. From the table we can see that a 0.5 TeV NLC has a reach at least double of that at RunIIa and still higher than that of RunIIb. The reach by the LHC is only slightly better than the 1 TeV NLC. ## 3 Kaluza-Klein States of SM Gauge Bosons In another brane configuration, the SM particles reside on a $`p`$-brane ($`p=\delta +3>3`$) whereas the gravity is in the rest $`(6p)`$ dimensional bulk. Within this $`p`$-brane the effect of gravity is negligible compared to gauge interactions. Inside the $`p`$-brane the chiral fermions are restricted to a 3-brane and the gauge bosons can propagate in the extra $`\delta `$ dimensions internal to the $`p`$-brane. Dienes et al. considered in this model of extra dimensions and showed that the gauge couplings can be unified at a scale lower than the usual GUT scale, due to the power running of the couplings. Therefore, the effective GUT scale is lowered, which is in contrast to the ADD model that the Planck scale is lowered. For collider phenomenology a specific model of this type was proposed by Pomarol et al. . It is a five-dimensional model with the fifth dimension compactified on a $`S^1/Z_2`$. In the four-dimensional point of view, a gauge boson $`V`$ that propagates in the fifth dimension is equivalent to a tower of Kaluza-Klein states $`V^{(n)}`$ with mass $`M_n=nM_c`$, where $`M_c=1/R`$ is the compactification scale and $`R`$ is radius of the fifth dimension. The resulting 4-D Lagrangian for charged-current and neutral-current interactions are, respectively, given by $`^{\mathrm{CC}}`$ $`=`$ $`{\displaystyle \frac{g^2v^2}{8}}[W_1^2+\mathrm{cos}^2\beta {\displaystyle }_{n=1}^{\mathrm{}}(W_1^{(n)})^2+2\sqrt{2}\mathrm{sin}^2\beta W_1{\displaystyle }_{n=1}^{\mathrm{}}W_1^{(n)}+2\mathrm{sin}^2\beta `$ (5) $`\times `$ $`\left({\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}W_1^{(n)}\right)^2]+{\displaystyle \frac{1}{2}}{\displaystyle }_{n=1}^{\mathrm{}}n^2M_c^2(W_1^{(n)})^2g(W_1^\mu +\sqrt{2}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}W_1^{(n)\mu })J_\mu ^1`$ $`+`$ $`(12)`$ $`^{\mathrm{NC}}`$ $`=`$ $`{\displaystyle \frac{gv^2}{8c_\theta ^2}}[Z^2+\mathrm{cos}^2\beta {\displaystyle }_{n=1}^{\mathrm{}}(Z^{(n)})^2+2\sqrt{2}\mathrm{sin}^2\beta Z{\displaystyle }_{n=1}^{\mathrm{}}Z^{(n)}+2\mathrm{sin}^2\beta `$ (6) $`\times `$ $`\left({\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}Z^{(n)}\right)^2+{\displaystyle \frac{1}{2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}n^2M_c^2[(Z^{(n)})^2+(A^{(n)})^2]`$ $``$ $`{\displaystyle \frac{e}{s_\theta c_\theta }}\left(Z^\mu +\sqrt{2}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}Z^{(n)\mu }\right)J_\mu ^Ze\left(A^\mu +\sqrt{2}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}A^{(n)\mu }\right)J_\mu ^{\mathrm{em}}`$ There are two important effects of these KK states on collider experiments. (i) Since the KK states have the same quantum numbers as their corresponding gauge bosons, it gives rise to mixing effects between the zeroth (the SM gauge boson) and the $`n`$th-modes of $`W`$ and $`Z`$ bosons. The zero mass of the photon is protected by the U(1)<sub>EM</sub> symmetry of the SM. (ii) If the energy is higher than the compactification scale $`M_c`$, real emissions or resonances of KK states of $`\gamma ,W,Z,g`$ bosons can be observed, otherwise enhancement of cross sections may be possible. In Eqs. (5) and (6) the first few terms of each imply mixings among $`V,V^{(1)},V^{(2)},\mathrm{}`$ ($`V=W,Z`$). The mixing between the SM gauge bosons with its Kaluza-Klein states modifies electroweak observables (similar to the mixing between the $`Z`$ and a $`Z^{}`$) via a series of mixing angles, which depend on the masses of $`Z^{(n)},n=0,1,\mathrm{}`$ and the angle $`\beta `$. The neutral boson at LEP is then the first mass eigenstate after mixing. The couplings of the $`Z^{(0)}`$ to fermions are modified through the mixing angles. The observables at LEPI can place strong constraints on the mixing, and thus on the compactification scale $`M_c`$. Similarly, the properties of the $`W`$ boson are also modified. The effects on electroweak precision measurements have been studied . Overall, the limit on $`M_c`$ using the precision data measurements is $`M_c\stackrel{>}{}3.33.8`$ TeV. The effects of Kaluza-Klein states of the SM gauge bosons also occur in high energy processes. If the available energy is higher than the compactification scale, real emissions or resonances of these Kaluza-Klein states can be observed. However, for the present collider energies and because the compactification scale is believed to be at least a few TeV, only indirect effects can be seen. We summarize a study of high energy processes when $`M_c`$ is higher than the present energy scale. The reduced amplitude for $`q\overline{q}\mathrm{}^+\mathrm{}^{}`$ or $`\mathrm{}^+\mathrm{}^{}q\overline{q}`$ is given by $$M_{\alpha \beta }^{eq}(s)=e^2\{\frac{Q_eQ_q}{s}+\frac{g_\alpha ^eg_\beta ^q}{s_\theta ^2c_\theta ^2}\frac{1}{sM_Z^2}(Q_eQ_q+\frac{g_\alpha ^eg_\beta ^q}{s_\theta ^2c_\theta ^2})\frac{\pi ^2}{3M_c^2}\}.$$ (7) where $`M_c^2s,|t|,|u|`$. The above formula is applicable to hadronic and leptonic cross sections at $`e^+e^{}`$ colliders and to Drell-Yan production at the Tevatron, and with a crossing to DIS at HERA. Similar expressions can be found for $`W`$ KK state exchanges and for dijet and $`t\overline{t}`$ production. In a global fit to $`\eta =\pi ^2/3M_c^2`$, Cheung and Landsberg include the following data sets: (i) LEPII hadronic and all leptonic production cross sections and angular distributions, (ii) Drell-Yan production at the Tevatron, (iii) NC and CC DIS scattering cross sections at HERA, (iv) Tevatron dijet production, and (v) Tevatron $`t\overline{t}`$ production. The resulting 95% C.L. limit on $`M_c`$ is $$M_c>4.4\mathrm{TeV}.$$ ## 4 Conclusions Physics in extra dimensions and phenomenology are extremely rich with advance in string theories because fundamental scales are now reachable within future collider experiments. We have briefly reviewed collider signatures of two interesting scenarios. The first model is due to the KK states of gravitons while the second one is due to KK states of gauge bosons. It turns out that diphoton, boson-pair, and fermion-pair production, as well as precision data measurements are useful in probing these two models. ## Acknowledgments I thank Wai-yee Keung and Greg Landsberg for parts of the work presented here. This research was supported in part by the DOE Grants No. DE-FG03-91ER40674.
warning/0003/astro-ph0003474.html
ar5iv
text
# Computer simulations of interferometric imaging with the VLT interferometer and the AMBER instrument ## 1 INTRODUCTION The Very Large Telescope Interferometer (VLTI) with its four 8.2 m unit telescopes (UTs) and three 1.8 m auxiliary telecopes (ATs) will certainly establish a new era of optical and infrared interferometric imaging within the next few years. With a maximum baseline of up to more than 200 m, the VLTI will allow the study of astrophysical key objects with unprecendented resolution opening up new vistas to a better understanding of their physics. The near-infrared focal plane instrument of the VLTI, the Astronomical MultiBEam Recombiner (AMBER), will operate between 1 and 2.5 $`\mu `$m and for up to three beams allowing the measurement of closure phases. In a second phase its wavelength coverage is planned to be extended to 0.6 $`\mu `$m. Objects as faint as $`K=20`$ mag are expected to be observable with AMBER when a bright reference star is available, and as faint as $`K=14`$ mag otherwise. Among the astrophysical key issues are, for instance, young stellar objects, active galactic nuclei and stars in late phases of stellar evolution. The simulation of a stellar object consists in principle of two components: (i) the calculation of an astrophysical model of the object, typically based on radiative transfer calculations predicting, e.g., its intensity distribution. To obtain a robust and non-ambiguous model, it is of particular importance to take diverse observational constraints into account, for instance the spectral energy distribution and visibilities. (ii) the determination of the interferometer’s response to this intensity signal, i.e. the simulation of light propagation in the atmosphere and the interferometer. Often, only one of the above parts is considered in full detail. The aim of this study is to combine both efforts and to present a computer simulation of the VLTI performance for observations of one object class, the dusty supergiants. For this purpose, we calculated a detailed radiative transfer model for one of its most outstanding representatives, the supergiant IRC +10 420, and carried out computer simulations of VLTI visibility measurements. The goal is to estimate how accurate visibilities can be measured with the VLTI in this particular but not untypical case, to discuss if the accuracy is sufficient to distinguish between different theoretical model predictions and to study on what the accuracy is dependent. ## 2 The supergiant IRC +10 420: Evolution on human timescales The star IRC +10 420 (= V 1302 Aql = IRAS 19244+1115) is an outstanding object for the study of stellar evolution. Its spectral type changed from F8 I$`{}_{}{}^{+}{}_{\mathrm{a}}{}^{}`$ in 1973 to mid-A today corresponding to an effective temperature increase of 1000-2000 K within only 25 yr. It is one of the brightest IRAS objects due to its very strong infrared excess by circumstellar dust and one of the warmest stellar OH maser sources known. Large mass-loss rates, typically of the order of several $`10^4`$ M/yr were determined by CO observations. IRC +10 420 is believed to be a massive luminous star currently being observed in its rapid transition from the red supergiant stage to the Wolf-Rayet phase. Its massive nature (initially $`20`$ to 40 M) can be concluded from its distance ($`d`$ = 3–5 kpc) and large wind velocity (40 km/s) ruling out alternative scenarios. IRC +10 420 is the only object observed until now in its transition to the Wolf-Rayet phase. The structure of the circumstellar environment of IRC +10 420 appears to be very complex, and scenarios proposed to explain the observed spectral features of IRC +10 420 include a rotating equatorial disk, bipolar outflows, and the infall of circumstellar material onto the star’s photosphere . Several infrared speckle and coronographic observations were conducted to study the dust shell of IRC +10 420. The most recent study reports the first diffraction-limited 73 mas bispectrum speckle interferometry of IRC +10 420 and presents the first radiative transfer calculations that model both the spectral energy distribution and the visibility of this key object. In the following we will briefly describe the main results and conclusions of this study which will serve as astrophysical input for the VLTI computer simulations presented in the next section. ### 2.1 Visibility measurements and bispectrum speckle interferometry Speckle interferograms of IRC +10 420 were obtained with the Russian 6 m telescope at the Special Astrophysical Observatory on June 13 and 14, 1998. The speckle data were recorded with our NICMOS-3 speckle camera through an interference filter with a centre wavelength of 2.11 $`\mu `$m and a bandwidth of 0.19 $`\mu `$m. The observational parameters were as follows: exposure time/frame 50 ms; number of frames 8400; 2.11 $`\mu `$m seeing (FWHM) $``$1$`.^{\prime \prime }`$0. A diffraction-limited image of IRC +10 420 with 73 mas resolution was reconstructed from the speckle interferograms using the bispectrum speckle interferometry method. The modulus of the object Fourier transform (visibility) was determined with the speckle interferometry method. Figure 1 (left panel) shows the reconstructed 2.11 $`\mu `$m visibility function of IRC +10 420. There is only marginal evidence for an elliptical visibility shape (position angle of the long axis $`130{}_{}{}^{}\pm 20^{}`$, axis ratio $`1.0`$ to 1.1). The visibility 0.6 at frequencies $`>4`$ cycles/arcsec shows that the stellar contribution to the total flux is $``$ 60% and the dust shell contribution is $``$ 40%. The Gauß fit FWHM diameter of the dust shell was determined to be $`d_{\mathrm{FWHM}}=(219\pm 30)`$ mas. By comparison, previous 3.8 m telescope K-band observations found a dust-shell flux contribution of $``$50% and $`d_{\mathrm{FWHM}}=216`$ mas. However, as will be shown later, a ring-like intensity distribution appears to be much better suited than the assumption of a Gaussian distribution whose corresponding FWHM diameter fit may give misleading sizes (see Sect. 2.2). The right panel of Fig. 1 displays the azimuthally averaged diffraction-limited images of IRC +10 420 and the unresolved star HIP 95447. The $`K`$-band visibility will strongly constrain radiative transfer calculations as shown in the next section. ### 2.2 Dust-shell models The spectral energy distrubion (SED) of IRC +10 420 with 9.7 and 18 $`\mu `$m silicate emission features is shown in Fig. 2. It corresponds to the ’1992’ data set used by Oudmaijer et al.) and combines VRI (October 1991), near-infrared (March and April 1992) and Kuiper Airborne Observatory (June 1991) photometry with the IRAS measurements and 1.3 mm data. Additionally, we included data for $`\lambda <0.55\mu `$m. In contrast to the near-infrared, the optical magnitudes have remained constant during the last twenty years within a tolerance of $`0.^\mathrm{m}1`$. IRC +10 420 is highly reddened due to an extinction of $`A_\mathrm{V}^{\mathrm{total}}7^\mathrm{m}`$ by the interstellar medium and the circumstellar shell. From polarization studies an interstellar extinction of $`A_\mathrm{V}6^\mathrm{m}`$ to $`7^\mathrm{m}`$ was estimated . Based on the strength of the diffuse interstellar bands, $`E(BV)=1.^\mathrm{m}4\pm 0.^\mathrm{m}5`$ can be inferred for the interstellar contribution compared to a total of $`E(BV)=2.^\mathrm{m}4`$. We will use an interstellar $`A_\mathrm{V}`$ of $`5^\mathrm{m}`$ adopting the method of Savage & Mathis with $`A_\mathrm{V}=3.1E(BV)`$. In order to model both the observed SED and $`2.11\mu `$m visibility, we performed radiative transfer calculations for dust shells assuming spherical symmetry. We used the code DUSTY which solves the spherical radiative transfer problem utilizing the self-similarity and scaling behaviour of IR emission from radiatively heated dust. To solve the radiative transfer problem including absorption, emission and scattering several properties of the central source and its surrounding envelope are required, viz. (i) the spectral shape of the central source’s radiation; (ii) the dust properties, i.e. the envelope’s chemical composition and grain size distribution as well as the dust temperature at the inner boundary; (iii) the relative thickness of the envelope, i.e. the ratio of outer to inner shell radius, and the density distribution; and (iv) the total optical depth at a given reference wavelength. The code has been expanded for the calculation of synthetic visibilities. We calculated various models of the dust shell of IRC +10 420 considering black bodies and model atmospheres as central sources of radiation, different silicates, grain-size and density distributions. We refer to Blöcker et al. for a full description of the model grid. The remaining fit parameters are the dust temperature, $`T_1`$, which determines the radius of the shell’s inner boundary, $`r_1`$, and the optical depth, $`\tau `$, at a given reference wavelength, $`\lambda _{\mathrm{ref}}`$. We refer to $`\lambda _{\mathrm{ref}}=0.55\mu `$m. Models were calculated for dust temperatures between 400 and 1000 K and optical depths between 1 and 12. Significantly larger values for $`\tau `$ lead to silicate features in absorption. The near-infrared visibility strongly constrains dust shell models since it is, e.g., a sensitive indicator of the grain size. Accordingly, high-resolution interferometry results provide essential ingredients for models of circumstellar dust-shells. In the instance of IRC +10 420 (assuming the central star to be a black body of $`T_{\mathrm{eff}}=7000`$ K), the silicate grain sizes, $`a`$, were found to be in accordance with a standard distribution function, $`n(a)`$$``$$`a^{3.5}`$, with $`a`$ ranging between $`a_{\mathrm{min}}`$ = 0.005 $`\mu `$m and $`a_{\mathrm{max}}`$ = 0.45 $`\mu `$m. However, the observed dust shell properties cannot be fitted by single-shell models but seem to require multiple components. At a certain distance we considered an enhancement over the assumed $`1/r^x`$ density distribution. The best model for both SED and visibility was found for a dust shell with a dust temperature of 1000 K at its inner radius of $`r_1=69R_{}`$. At a distance of $`r=308R_{}`$ ($`Y=r/r_1=4.5`$) the density was enhanced by a factor of $`A=40`$ and and its density exponent was changed from $`x=2`$ to $`x=1.7`$. These fits for SED and 2.11 $`\mu `$m visibility are shown in Fig. 2. The various flux contributions at 2.11 $`\mu `$m are 62.2% stellar light, 26.1% scattered radiation and 10.7% dust emission (see Fig. 3), i.e. the radiation emitted by the circumstellar shell itself consists of 71% scattered radiation and 29% direct dust emission. The shell’s model intensity distribution is shown in Fig. 3 and was found to be ring-like. This appears to be typical for optically thin shells (here $`\tau _{0.55\mu \mathrm{m}}=7`$, $`\tau _{2.11\mu \mathrm{m}}=0.55`$) showing limb-brightened dust-condensation zones. Accordingly, the interpretation of the observational data by FWHM Gauß diameters may give misleading results. The ring diameter is equal to the inner diameter of the hot shell ($`69`$ mas), and the diameter of the central star amounts to $`1`$ mas. The bolometric flux, $`F_{\mathrm{bol}}`$, is $`8.1710^{10}`$ Wm<sup>-2</sup> corresponding to a central-star luminosity of $`L/L_{}=\mathrm{25\hspace{0.17em}462}(d/\mathrm{kpc})^2`$. This two-component model can be interpreted in terms of a termination of an enhanced mass-loss phase roughly 90 yr (for $`d=5`$ kpc) ago. The assumption that IRC +10 420 had passed through a superwind phase in its recent history is in line with its evolutionary status of an object in transition from the Red-Supergiant to the Wolf-Rayet phase. The mass-loss rates of the components can be determined to be $`\dot{M}_1=\mathrm{7.0\; 10}^5`$$`M_{}/\mathrm{yr}`$ and $`\dot{M}_2=\mathrm{1.1\; 10}^3`$$`M_{}/\mathrm{yr}`$. ## 3 Interferometry with the VLTI and the AMBER instrument In the previous section, the spatial intensity profile of the dusty supergiant IRC +10 420 (see Fig. 3) was derived by means of radiative transfer models and their comparison with photometric and interferometric observations. This $`2.11\mu `$m intensity profile will serve as object intensity profile in the simulation of monochromatic VLTI observations. The next steps are the simulation of light propagation from the object to the detector (through atmosphere, telescopes, and the AMBER wide-field mode instrument), simulation of photon noise and detector read-out noise, and finally data processing of the interferograms. A schematic view of Michelson interferometry with the VLTI and the AMBER camera is given in Fig. 4. ### 3.1 Computer simulation of interferometric imaging Fig. 5 shows a flow chart of our simulation of interferometric imaging with the VLTI (ATs, or UTs with adaptive optics) and the AMBER camera in the wide-field mode (i.e. without fiber optics spatial filtering). In a first step, an array of Gaussian distributed random numbers is generated and convolved with the correlation function of the atmospheric refraction index variations in order to generate wavefronts degraded by atmospheric turbulence. The typical size of the atmospheric turbulence cells is given by the Fried parameter $`r_0`$. After the simulation of the entrance pupil, the next step incorporates the tip-tilt correction but we allow for a residual tip-tilt error $`\delta _{\mathrm{tt}}`$. In the next step a typical Michelson output pupil is simulated (pupil reconfiguration). The output pupil is chosen such that (i) in the optical transfer function the off-axis peaks are separated from the central peak, and (ii) the interferograms are sampled with the smallest number of pixels to assure the lowest influence of detector noise. The following step includes light propagation through the beam combiner lens to the focal plane. The squared modulus of the Fourier transform of the complex amplitude in front of the beam combiner lens yields the intensity distribution of a Michelson interferogram of a point source. In the next step the required object intensity distribution is simulated (given here by the intensity distribution of IRC +10 420) to obtain the Michelson interferogram of the object: The Fourier transformation of the object intensity distribution, calculated at those spatial frequencies covered by the the simulated interferometer baseline vector, is multiplied with the off-axis peaks of the transfer function of the generated Michelson interferogram. Finally, Poisson photon-noise and detector read-out noise is injected to the interferograms. The noise level depends, among other parameters, on the number of detectable photons, the total optical throughput of the interferometer and the quantum efficiency of the detector. Details are shown in Table 1. ### 3.2 Visibility data for IRC +10 420 We performed simulations of IRC +10 420 AT visibility observations in the AMBER wide-field mode and studied the influence of various observational parameters on the visibility accuracy. Visibility error bars were, for example, obtained for the following observational parameters: different seeing during the observation of object and reference star (Fried parameters $`r_{0,\mathrm{object}}`$=2.4 m, $`r_{0,\mathrm{ref}.}`$=2.5 m), different residual tip-tilt error ($`\delta _{\mathrm{tt},\mathrm{object}}`$=2% of the Airy disk diameter, $`\delta _{\mathrm{tt},\mathrm{ref}.}`$=0.1%), and object brightness ($`K_{\mathrm{object}}`$=3.5 mag and 11 mag, $`K_{\mathrm{ref}.}`$=3.5 mag). In the computer experiments (a)-(c), object and reference star were assumed to have the same brightness ($`K`$=3.5 mag, see Table 1), in experiment (d) fainter objects ($`K`$=11 mag) were simulated as well. Fig. 6 shows the results of the simulations (a) to (c) based on the $`A=40`$ intensity profile (see Fig. 3) for baselines of 50 and 100 m together with the model predictions for IRC +10 420. To obtain error bars each simulation was repeated three times. The insetted table lists the parameters of the simulations. Simulation (a) based on 2400 interferograms represents the ideal case of excellent seeing conditions ($`r_{0,\mathrm{object}}`$=2.5 m, $`r_{0,\mathrm{ref}.}`$=2.5 m) and an almost perfect tip-tilt correction (residual tip-tilt error $`\delta _{\mathrm{tt}}=0.1`$%). The visibility error $`\mathrm{\Delta }V`$ amounts to $`\pm 0.0029`$. Simulation (b) illustrates the influence of a larger residual object tip-tilt error ($`\delta _{\mathrm{tt}}`$=2%) leading to $`\mathrm{\Delta }V\pm 0.0036`$. Simulation (c) shows the impact of different seeing conditions for object and reference star. In the last simulation (d) of this series (shown only in the insetted table of Fig. 6) an IRC +10 420-like intensity distribution is assumed but the simulated $`K`$-magnitude is 11 mag (instead of 3.5 mag). Although the photon number decreases to $`N=4.4110^3`$ (i.e. by a factor of 1000, see Table 1), the visibility accuracy is still high, i.e. $`\mathrm{\Delta }V\pm 0.0036`$. ## 4 Conclusions We have presented computer simulations of interferometric imaging with the VLT interferometer and the AMBER instrument in the wide-field mode. These simulations include both the astrophysical modelling of a stellar object by radiative transfer calculations and the simulation of light propagation from the object to the detector and simulation of photon noise and detector read-out noise. We focussed on stars in late stages of stellar evolution and examplarily studied one of its most outstanding representatives, the dusty supergiant IRC +10 420. The model intensity distribution of this key object, obtained by radiative transfer calculations, served as astrophyiscal input for the VLTI/AMBER simulations. The results of these simulations show the dependence of the visibility error bar on various observational parameters. With these simulations at hand one can immediately see under which conditions the visibility data quality would allow us to discriminate between different model assumptions (e.g. the size of the superwind amplitude). Inspection of Fig. 6 shows that in all studied cases the observations will give clear preference to one particular model. Therefore, observations with VLTI will certainly be well suited to gain deeper insight into the physics of dusty supergiants. ## ACKNOWLEDGMENTS The bispectrum speckle observations were made with the SAO 6 m telescope operated by the Special Astrophysical Observatory, Russia. The radiative-transfer calculations are based on the code DUSTY developed by Ž. Ivezić, M. Nenkova and M. Elitzur.
warning/0003/hep-ph0003229.html
ar5iv
text
# Untitled Document Higgs induced light leptoquark-diquark mixing and proton decay Uma Mahanta Mehta Research Institute Chhatnag Road, Jhusi Allahabad-211019, India email:mahanta@mri.ernet.in Abstract In low energy phenomenology to avoid the strong constraints of proton decay it is usually assumed that light ($``$ 250 Gev) leptoquarks couple only to quark-lepton pairs and light diquarks couple only to quark pairs. In this paper we present two specific examples where the higgs induced mixing between leptoquarks and diquarks through trilinear interaction terms reintroduces the troublesome couplings and gives rise to proton decay. The bound on the unknown parameters of this scenario that arise from proton life time has also been derived. Leptoquarks (LQ) and diquarks (DQ) are colored scalar or vector particles that carry baryon numbers of $`\pm \frac{1}{3}`$ and $`\pm \frac{2}{3}`$ respectively. They occur naturally in many extensions of the standard model (SM) e.g. superstring inspired grand unified models based on $`E(6)`$ , technicolor models and composite models of quarks and leptons. If LQs and DQs couple both to quark-lepton pairs and quark pairs then they lead to too rapid proton decay. To be consistent with the proton life time such LQs and DQs must have a mass of the order of $`(10^{12}10^{15})`$ Gev. To avoid this strong constraint and to make them relevant for low energy phenomenology below 1 Tev it is usually assumed that LQs couple to quark-lepton pairs but not to quark pairs and DQs couple to quark pairs but not to quark-lepton pairs. However in the presence of the SM higgs doublet $`\varphi `$ the low energy effective Lagrangian will also contain a trilinear Higgs-LQ-DQ interaction term. After EW symmetry breaking (EWSB) this interaction term will induce a mixing between the LQ and the DQ which reintroduces the troublesome Yukawa couplings for LQ and DQ. If the physical LQ and DQ do not have the same mass then they lead to proton decay. There can be several different kinds \[1-3\] of LQ’s and DQ’s. However in this report we shall not go into a detailed discussion of every LQ-DQ mixing that can possibly occur. Rather we shall give two give specific examples of Higgs-LQ-DQ interaction term and the mixing between the LQ and DQ that takes place after EWSB. For one particular case we have also calculated the mixing angle in terms of the dimensional coupling that measures the strength of the Higgs-LQ-DQ interaction term. Finally we have estimated the proton decay rate in terms of the unknown parameters of our scenario and have derived the constraint that the parameters of the Lagrangian must satisfy in order to be consistent with proton lifetime. Consider the SM to be extended by a light ($``$ 250 Gev ) chiral leptoquark D and a light chiral diquark S with the following assignments under $`SU(3)_c\times SU(2)_l\times U(1)_y`$: $`D(3^{},2,\frac{1}{6})`$ and $`S(3^{},1,\frac{1}{3})`$. The low energy effective Lagrangian of this extended scenario will contain besides the SM Lagranigian all possible renormalizable and gauge invariant interaction terms between D, S and the SM fields. Of particular importance for this work are the Yukawa like couplings of D and S to the SM fermions given by the following Lagrangian $$\begin{array}{ccc}\hfill L_y& =g_R\overline{l}_Ld_RD+g^{}ϵ_{ijk}\overline{d}_{Ri}u_{Rj}^cS_k+h.c.\hfill & \\ & =g_R(\overline{\nu }_LD_1+\overline{e}_LD_2)d_R+g_R^{}ϵ_{ijk}\overline{d}_{Ri}u_{Rj}^cS_k+h.c.\hfill & (1)\hfill \end{array}$$ Here $`D_1`$ and $`D_2`$ are the isospin up and down components of D. i, j, k are the color indices. Note first that the leptoquark D couples only to quark-lepton pair and the diquark S couples only to quark pair which is required so that they do not lead to proton decay. Second both interaction terms are invariant under the SM gauge group $`SU(3)_c\times SU(2)_l\times U(1)_y`$. Finally the couplings of D and S are both chiral in nature i.e. they couple to quark fields of a particular chirality only. The Yukawa Lagrangian $`L_y`$ also implies that D and S should carry baryon numbers of $`\frac{1}{3}`$ and $`\frac{2}{3}`$ respectively so that $`L_y`$ conserves baryon number. This kind of structure of the Yukawa couplings could arise depending on the gauge symmetry of the high energy theory and its chiral matter representation. Besides the Yukawa couplings the low energy Lagrangian for describing physics much below the compositeness scale or grand unified scale $`\mathrm{\Lambda }`$ must also contain all possible renormalizable and gauge invariant interaction terms between $`\varphi `$, D and S. In principle such interaction terms cannot be neglected. They could arise from the scalar potential of the high energy theory after the gauge symmetry of the high energy theory breaks down into $`SU(3)_c\times SU(2)_l\times U(1)_y`$. The gauge invariant trilinear interaction term between $`\varphi `$, D and S is given by the Lagrangian $$L_s=k_1(D^+\varphi _c)S+h.c.=k_1\frac{v+h}{\sqrt{2}}D_1^+S+h.c.$$ $`(2)`$ Here $`\varphi _c`$ is the charge conjugated Higgs doublet that gives mass to the up quarks in SM. The unknown mixing parameter $`k_1`$ carries the dimension of mass. After EW symmetry breaking this interaction term will lead to mixing between $`D_1`$ and S. Note that $`D_1`$ and S carries the same charge and color assignments which is necessary so that they could mix after EWSB. The above higgs-LQ-DQ interaction term violates both baryon number and lepton number by one unit ($`\delta B=\delta L=1`$). Here we are implicitly assuming that the breaking of baryon number symmetry is first communicated to the scalar sector of the high energy theory which in turn communicates it to other sectors e.g the Yukawa sector. Recall that in the SM baryon number conservation is realized as an accidental global symmetry . By that we mean that given the particle content and the gauge group $`SU(3)_c\times SU(2)_l\times U(1)_y`$ of the SM it is not possible to write a renormalizable and gauge invariant interaction term that violates baryon number. No adhoc global symmetry is required to explain the near abscence of proton decay. However as we have seen above that if we keep the gauge group the same but allow additional particles like color triplet scalars then it is possible to write down a renormalizable and gauge invariant interaction term that violates baryon number. We shall now show that the mixing between $`D_1`$ and S induced by $`\varphi `$ can lead to proton decay in the “$`\nu `$ \+ any channel” where any refers to a positively charged non-strange meson. Consider the full scalar potential involving D and S $$\begin{array}{ccc}\hfill V(D,S)& =\mu _1^2D^+D+\mu _2^2S^+S+\lambda _1(D^+D)^2+\lambda _2(S^+S)^2\hfill & \\ & +\lambda _1^{}(D^+D)(\varphi ^+\varphi )+\lambda _2^{}(S^+S)(\varphi ^+\varphi )+(k_1D^+\varphi _cS+h.c.)\hfill & (3)\hfill \end{array}$$ The quartic scalar interactions are not relevant to our present work. $`\mu _1^2`$ and $`\mu _2^2`$ are the mass parameters associated with the gauge eigenstates D and S. After EWSB the trilinear interaction term between $`\varphi `$, D and S induces a mixing between $`D_1`$ and S. It can be shown that the eigenvalues of the mass squared matrix $`M_{D_1S}^2`$ are given by $$\lambda _+=M_{D_1}^2=\frac{1}{2}(\mu _1^2+\mu _2^2)+\frac{1}{2}\sqrt{ϵ^2+2k_1^2v^2}.$$ $`(4a)`$ and $$\lambda _{}=M_S^2=\frac{1}{2}(\mu _1^2+\mu _2^2)\frac{1}{2}\sqrt{ϵ^2+2k_1^2v^2}.$$ $`(4b)`$ where $`ϵ=\mu _1^2\mu _2^2`$. The physical states correponding to $`\lambda _\pm `$ are given by $`D_1^{}=D_1\mathrm{cos}\theta S\mathrm{sin}\theta `$ and $`S^{}=D_1\mathrm{sin}\theta +S\mathrm{cos}\theta `$. Here $`\mathrm{cos}\theta =\frac{\sqrt{2}k_1v}{[2k_1^2v^2+(\sqrt{ϵ^2+2k_1^2v^2}ϵ)^2]^{\frac{1}{2}}}`$. The Yukawa couplings of the LQ and the DQ written in terms of mass eigenstates are given by $$\begin{array}{ccc}\hfill L_y& =g_R[\overline{\nu }_L(D_1^{}\mathrm{cos}\theta +S^{}\mathrm{sin}\theta +\overline{e}_LD_2]d_R\hfill & \\ & +ϵ_{ijk}g_R^{}\overline{d}_{Ri}u_{Rj}^c[S_k^{}\mathrm{cos}\theta D_{1k}^{}\mathrm{sin}\theta +h.c.]\hfill & (5)\hfill \end{array}$$ . The above Yukawa Lagrangian does lead to proton decay in the “$`\nu `$ \+ any” channel via the exchange of D and S particles. The effective four fermion Lagrangian for this decay is given by $$L_{eff}=g_Rg_R^{}\mathrm{sin}\theta \mathrm{cos}\theta ϵ_{ijk}(\overline{\nu }_Ld_{Rk})(\overline{u}_{Rj}^cd_{Ri})(\frac{1}{M_S^2}\frac{1}{M_D^2})+h.c.$$ $`(6)`$ This is the effective Lagrangian at a scale $`M^2=M_D^2M_S^2(250Gev)^2`$. In order to use it for proton decay it has to be renormalized down to a scale $`\mu ^2=(1Gev)^2`$ under the unbroken QCD and EM interactions. The EM corrections are small because the coupling itself is small. The QCD corrections are not that large either because $`\mathrm{ln}\frac{M^2}{\mu ^2}`$ is not large in our case. It can be shown that the proton decay rate arising from the above effective Lagrangian is given by $$\mathrm{\Gamma }(p\nu _e+any)\frac{1}{17\pi ^2R^3}G_{eff}^2m_q^2A^2(\frac{M}{\mu })$$ $`(7)`$ . Here $`A(\frac{M}{\mu })`$ includes the effects of renormalization group evolution from M to $`\mu `$. We shall assume it to be of order one. $`R=\frac{3}{4}`$ fm, $`m_q\frac{1}{3}`$ Gev and $`G_{eff}=\frac{1}{2\sqrt{2}}g_Rg_R^{}\mathrm{sin}\theta \mathrm{cos}\theta (\frac{1}{M_S^2}\frac{1}{M_D^2})`$. The present lower bound on $`\tau (p\nu +any)`$ is 25$`\times 10^{30}`$ yr . It is clear from the expression of $`G_{eff}`$ that the tight constraint of proton lifetime can be satisfied either with $`\mathrm{sin}\theta 1`$ or $`\mathrm{cos}\theta 1`$. It can be shown that for $`\mathrm{sin}\theta 1`$ we require $`k_1v\sqrt{2}ϵ`$. On the other hand for $`\mathrm{cos}\theta 1`$ we need $`\sqrt{2}k_1v(\sqrt{ϵ^2+2k_1^2v^2}ϵ)`$. For $`M_S=200`$ Gev and $`M_D=300`$ Gev the product combination $`g_Rg_R^{}\mathrm{sin}\theta \mathrm{cos}\theta `$ must be less than $`3\times 10^{26}`$ in order that the decay rate in eqn(7) is consistent with the proton lifetime. If we assume that the Yukawa couplings $`g_R`$ and $`g_R^{}`$ are of the order of .1, then for $`\mathrm{sin}\theta 1`$ we must adjust the parameters of the Lagrangian so as to satisfy $`k_1v<4\times 10^{24}ϵ`$. On the other hand for $`\mathrm{cos}\theta 1`$ the parameters must be adjusted so as to satisfy $`\sqrt{2}k_1v<3\times 10^{24}(\sqrt{ϵ^2+2k_1^2v^2}ϵ)`$. These constraints may not be stable against radiative corrections requiring extreme fine tuning to each order in loop expansion. We would like to reiterate that there can be several different species of LQ’s as well as DQ’s. However the mixing between $`D_1`$ and S and hence proton decay occurs only for specific $`SU(3)_c\times SU(2)_l\times U(1)_y`$ assignments of the LQ and the DQ. Unless these assignments are achieved there is no mixing between the LQ and DQ and hence no contribution to proton decay. Having shown that higgs induced mixing between the doublet leptoquark D and the singlet diquark S does lead to proton decay we shall now present another concrete example where the same phenomenon also takes place. An EW triplet diquark $`T_{ak}`$ (a is the $`SU(2)_l`$ index and k the color index) can also mix with the doublet leptoquark D and contribute to both proton decay and neutron decay. Let the $`SU(3)_c\times SU(2)_l\times U(1)_y`$ assignments of $`T_{ak}`$ be given by (3, 3, $`\frac{1}{3}`$). The triplet diquark $`T_{ak}`$ can couple to LH quark pair according to the following Lagrangian: $$\begin{array}{ccc}\hfill L_y^{}& =g_R^{\prime \prime }ϵ_{ijk}\overline{q}_{Li}^c\tau _ai\tau _2T_{ak}+h.c.\hfill & \\ & =g_R^{\prime \prime }ϵ_{ijk}[\sqrt{2}\overline{u}_{Li}^cu_{Lj}T_k+\sqrt{2}\overline{d}_{Li}^cd_{Lj}T_{+k}+(\overline{u}_{Li}^cd_{Lj}+\overline{d}_{Li}^cu_{Lj})T_{0k}]\hfill & (6)\hfill \end{array}$$ where $`T_{+k}=\frac{T_{1k}+iT_{2k}}{\sqrt{2}}`$, $`T_k=\frac{T_{1k}iT_{2k}}{\sqrt{2}}`$ and $`T_{0k}=T_{3k}`$. $`T_{+k}`$ and $`T_k`$ carry electromagnetic charges of $`\frac{2}{3}`$ and $`\frac{4}{3}`$ units respectively. Note that $`T_k`$ is not the antiparticle of $`T_{+k}`$ since they carry different electromagnetic charges. The gauge invariant Higgs-LQ-DQ interaction term in this case will be given by $$\begin{array}{ccc}\hfill L_s^{}& =k_2(D^+\tau _a\varphi _c)T_a^{}+h.c.\hfill & \\ & =k_2\frac{v+h}{\sqrt{2}}[D_2^+T_+^{}+D_1^+T_0^{}+h.c.]\hfill & (7)\hfill \end{array}$$ The mixing between $`D_1`$ and $`T_0^{}`$leads to $`p\pi ^+\nu `$ and the mixing between $`D_2`$ and $`T_+^{}`$ leads to $`n\pi ^+e^{}`$ both of which violate baryon number. The low energy effective Lagrangian will also contain trilinear Higgs-LQ-LQ and Higgs-DQ-DQ interaction terms. Afetr EWSB these terms give rise to mixing between different multiplets of LQ and DQ. Such interaction terms do not violate baryon number but they could violate lepton number. Of particular importance is the Higgs-LQ-LQ ineraction for a pair of chiral leptoquarks belonging to different weak SU(2) multiplets. The mixing between two different LQ multiplets will give rise to helicity unsuppressed contribution to $`\pi ^{}e^{}\overline{\nu }_e`$ The helicity suppressed SM contribution to $`\pi ^{}e^{}\overline{\nu }_e`$ is in excellent agreement with the experimental data. Therefore any helicity unsuppressed contribution arising from new physics must be strongly constrained. This leads to stringent bounds on the mixing parameter between the two LQ multiplets. The mixing between different LQ multiplets can also lead to majorana mass matrix for neutrinos. A complete discussion of these topics can be found in Ref. and will not be taken up here. To conclude in this report we have shown that the usual assumption of low energy phenomenology that LQ’s couple only to q-l pairs and DQ’s couple only to quark pairs is not sufficient to stabilize the proton. We have given two specific examples where the Higgs-LQ-DQ interaction term induces a mixing between the LQ and DQ after EWSB. This mixing reintroduces the troublesome couplings for the LQ and the DQ and leads to proton decay. We find that in order to be consistent with the bound on proton lifetime the parameters of the Lagrangian must either satisfy $`k_1v<4\times 10^{24}ϵ`$ or $`\sqrt{2}k_1v<3\times 10^{24}(\sqrt{ϵ^2+2k_1^2v^2}ϵ)`$. In deriving these constraints we have assumed that $`g_Rg_R^{}.1`$, $`M_{D_1}=300`$ Gev and $`M_S`$=200 Gev. These constraints may not be stable against radiative corrections requiring extreme fine tuning of the parameters to each order in loop expansion. This naturalness problem can perhaps be resloved by means of some new symmetry that remains unbroken in the low energy theory and prevents the Higgs-LQ-DQ interaction term from occuring in the effective Lagrangian. References 1. W. Buchmuller Acta Phys. Austriaca, Suppl 27, 517 (1985); W. Buchmuller and D. Wyler, Phys. Lett. B 177, 377 (1986). 2. E. Farhi and L. Susskind, Phys. Rep. 74, 277 (1981); K. Lane and M. V. Ramanna, Phys. Rev. D 44, 2678 (1991). 3. J. Hewett and T. Rizzo, Phys. Rep. 183, 193 (1989). 4. A. Cohen, A. Nelson and D. Kaplan, Phys. Lett. B 388, 588 (1996). 5. Review of Particle Properties Euro. Phy. Jour. C3, 50 (1998). 6. The values of $`M_S`$ and $`M_D`$ assumed by us are consistent with the latest bounds on LQ and DQ masses: Euro. Phys. Jour. C3, 260 (1998); B. Abbot et al. (DO Collaboration), Phys. Rev. Lett. 80, 2051 (1998); F. Abe et al. (CDF Collaboration), Phys. Rev. D 55, R5263 (1997). 7. M. Hirsch et al., Phys. Rev. D 54, 4207(1996); Phys. Lett. B 378, 17 (1996); U. Mahanta, MRI-PHY/990930, hep-ph/9909518.
warning/0003/astro-ph0003368.html
ar5iv
text
# Angular Momentum Transport in Particle and Fluid Disks ## 1 Introduction When turbulence in unmagnetized Keplerian disks is driven by thermal convection, angular momentum appears to be transported inwards (Ryu & Goodman 1992; Kley, Papaloizou, & Lin 1993; Stone & Balbus 1996; Cabot 1996). If generically true, this implies that convection cannot provide the “enhanced” angular momentum transport (relative to collisional viscosity) needed to account for observations of accreting systems in astrophysics. Despite appearances, inward transport does not contradict the fact that an isolated Keplerian disk relaxes to its minimum energy state by transporting angular momentum outwards (Lynden-Bell & Pringle 1974). The reason is that simulated convective disks are not completely isolated. They tap a source of free energy other than their intrinsic differential rotation: an externally imposed, superadiabatic entropy gradient. A more mechanistic understanding of the behavior of turbulent fluid disks can be obtained by considering disks composed of discrete macroscopic particles. Transport properties and stability criteria obtained under one picture often carry over to the other, to within dimensionless factors of order unity. The fluid/particle dualism has been employed with considerable success in studies of planetary rings (see, e.g., Goldreich & Tremaine 1982) and galactic structure (see, e.g., Chapter 6 of Binney & Tremaine 1987). Indeed, one-to-one correspondences can be established between terms in the Reynolds stress tensor equation describing turbulent fluid disks, and terms in the Boltzmann stress tensor equation describing particle disks. This paper explores the transport properties of the latter system, with the aim of shedding light on the former. We demonstrate how simple particle disk models can reproduce many of the transport phenomena found in numerical simulations of vertically or radially convective fluid disks. The particle viewpoint also provides insight into the stability properties of isolated, unmagnetized, fluid disks. An analysis similar to ours was carried out by Kumar, Narayan, & Loeb (1995); this is discussed in §5. In §2, we write down the Boltzmann stress tensor equation governing particle disks and discuss its intimate relationship with the Reynolds stress tensor equation governing hydrodynamic turbulence in a fluid disk. In §3, we prove a general theorem that relates the direction of angular momentum transport to the behavior of the velocity dispersion tensor in the presence of external driving. Section 4 is the most important of this paper; there we consider a simple model for the collision term in the Boltzmann equation that allows one to easily solve for the stress tensor of the disk and, we believe, capture much of the relevant physics regarding both stability and the direction of transport. In §5, we summarize and discuss our results. ## 2 Stress Tensor Equations for Particle Disks The second moments of the Boltzmann equation describe the evolution of the viscous stress tensor. For an axisymmetric disk, in cylindrical coordinates, these read (see, e.g., Borderies, Goldreich, & Tremaine 1983): $$_tp_{rr}4\mathrm{\Omega }p_{r\varphi }=\left(_tp_{rr}\right)_c+\dot{H}_{rr}$$ (1) $$_tp_{\varphi \varphi }+\frac{\kappa ^2}{\mathrm{\Omega }}p_{r\varphi }=\left(_tp_{\varphi \varphi }\right)_c+\dot{H}_{\varphi \varphi }$$ (2) $$_tp_{zz}=\left(_tp_{zz}\right)_c+\dot{H}_{zz}$$ (3) $$_tp_{r\varphi }+\frac{\kappa ^2}{2\mathrm{\Omega }}p_{rr}2\mathrm{\Omega }p_{\varphi \varphi }=\left(_tp_{r\varphi }\right)_c,$$ (4) where we have neglected third moments. In equations (1)–(4), $`\mathrm{\Omega }`$ is the angular frequency, $`\kappa ^2=r^3d(\mathrm{\Omega }r^2)^2/dr`$ is the square of the epicyclic frequency, and $`p_{ij}`$ are the components of the stress tensor. The terms $`\left(_tp_{ij}\right)_c`$ describe changes in velocity dispersions due to interparticle collisions. Collisions perform three tasks: they (1) extract energy from the mean shear into random motions, (2) transfer energy in random motions associated with one direction to those associated with another, and (3) convert kinetic energy into heat. The terms $`\dot{H}_{ij}`$ are inserted to simulate external stirring of the disk, with $`\dot{H}_{r\varphi }=0`$ so as not to explicitly induce angular momentum transport. ### 2.1 Boltzmann vs. Reynolds Since the fluid equations are derived from the Boltzmann equation it is not surprising that hydrodynamic turbulence in fluid disks satisfies energy conservation equations very similar to equations (1)–(4). The Reynolds stress equation can be obtained by taking the Navier-Stokes equation, decomposing the velocity and pressure into mean and fluctuating components, and multiplying the entire equation by the velocity. This yields (e.g., Speziale 1991; Kato & Yoshizawa 1997): $$_tT_{ij}=T_{ik}_kU_jT_{jk}_kU_i+\pi _{ij}ϵ_{ij}_kC_{ijk},$$ (5) where $`U`$ is the mean velocity of the fluid, $`u`$ is the fluctuating velocity, $`T_{ij}=u_iu_j`$ is the stress tensor, and $``$ denotes an ensemble average. The quantity $`C_{ijk}`$ is a sum of turbulent and viscous fluxes, $$\pi _{ij}=u_i_jp+u_j_ip$$ (6) and $$ϵ_{ij}=2\nu _ku_i_ku_j,$$ (7) where $`p`$ is the fluctuating gas pressure and $`\nu `$ is the kinematic viscosity. The first two terms on the right hand side of equation (5) describe how turbulence extracts energy from the shear in the mean flow. The terms $`\pi _{ij}`$ are known as the “pressure-strain” correlation in the fluid dynamics literature;<sup>4</sup><sup>4</sup>4To be precise, our $`\pi _{ij}`$ is related to the usual pressure-strain correlation tensor by a total divergence, which we have absorbed into $`C_{ijk}`$. the diagonal terms of $`\pi _{ij}`$ are simply the work done by fluctuating pressure forces associated with the turbulence; among their other tasks, these terms attempt to enforce isotropy of the turbulence by transferring energy from velocity fluctuations associated with one direction to those associated with another. The terms $`ϵ_{ij}`$ represent the viscous dissipation of energy on small scales (note that $`ϵ_{ii}`$ is positive definite). It is straightforward to evaluate the above expression for $`_tT_{ij}`$ in cylindrical coordinates. We neglect the flux divergence<sup>5</sup><sup>5</sup>5This assumes “homogeneous” turbulence, i.e., that the flux varies on a length scale $`r`$, much greater than the vertical scale height of the disk. to obtain $$_tT_{rr}4\mathrm{\Omega }T_{r\varphi }=\pi _{rr}ϵ_{rr}$$ (8) $$_tT_{\varphi \varphi }+\frac{\kappa ^2}{\mathrm{\Omega }}T_{r\varphi }=\pi _{\varphi \varphi }ϵ_{\varphi \varphi }$$ (9) $$_tT_{zz}=\pi _{zz}ϵ_{zz}$$ (10) $$_tT_{r\varphi }+\frac{\kappa ^2}{2\mathrm{\Omega }}T_{rr}2\mathrm{\Omega }T_{\varphi \varphi }=\pi _{r\varphi }ϵ_{r\varphi }.$$ (11) Equations (8)–(10) are identical to the energy equations given by, e.g., Balbus & Hawley (1998; see their equations 58–60). Equations (1)–(4) for the evolution of the stress tensor in a particle disk are very similar to equations (8)–(11) for the evolution of the stress tensor in a fluid disk. In particular, the rotational profile of the disk enters identically in the two cases; both $`\mathrm{\Omega }`$ and $`\kappa ^2`$ couple to the stress tensor components in exactly the same way for particle and fluid disks. Moreover, the “source” and “sink” terms in the energy equations are analogous; the collision term, $`\left(_tp_{ij}\right)_c`$, in the Boltzmann equation plays the same role as $`\pi _{ij}ϵ_{ij}`$ in the Reynolds stress equations. This comparison indicates that the dynamical role of epicycles is the same in both fluid and particle disks, and that any difference in the angular momentum transport properties of the two disks can only be due to differences in the sources and sinks in the energy equation. We demonstrate this formally in the next section by proving a theorem that relates the direction of angular momentum transport to the properties of the collisional and stirring terms. This theorem is very useful for interpreting the model problem of §4. ## 3 The “Long Axis Must Get Longer” Theorem Theorem: For thin axisymmetric disks in which $`d\mathrm{\Omega }/dr<0`$, steady inward transport of angular momentum requires that the net tendency of collisions and external stirring be to further lengthen the principal equatorial long axis of the velocity ellipsoid and to further shorten the principal equatorial short axis. This theorem applies to fluid disks as well as to particle disks so long as the interparticle collision term, $`(_tp_{ij})_c`$, is replaced by $`\pi _{ij}ϵ_{ij}`$. Proof: We transform equations (1)–(4) to the principal axis basis of the velocity ellipsoid. Let $`p_{11}`$ and $`p_{22}`$ be the greater and lesser of the two components of the stress tensor, respectively, in the equatorial plane: $`p_{11}`$ $`=`$ $`[p_{rr}+p_{\varphi \varphi }+\sqrt{(p_{rr}p_{\varphi \varphi })^2+4p_{r\varphi }^2}]/\mathrm{\hspace{0.17em}2}`$ (12) $`p_{22}`$ $`=`$ $`[p_{rr}+p_{\varphi \varphi }\sqrt{(p_{rr}p_{\varphi \varphi })^2+4p_{r\varphi }^2}]/\mathrm{\hspace{0.17em}2},`$ (13) where we shall always take the positive square root. By symmetry about the midplane, $`p_{33}=p_{zz}`$. The orientation of the velocity ellipsoid is given by $`\delta `$, the angle between the longer equatorial axis and the radial direction: $`\mathrm{tan}\delta `$ $``$ $`{\displaystyle \frac{p_{r\varphi }}{p_{11}p_{\varphi \varphi }}}`$ (14) $`=`$ $`{\displaystyle \frac{2p_{r\varphi }}{(p_{rr}p_{\varphi \varphi })+\sqrt{(p_{rr}p_{\varphi \varphi })^2+4p_{r\varphi }^2}}},`$ valid for $`|\delta |<\pi /2`$ and nonisotropic $`p_{ij}`$.<sup>6</sup><sup>6</sup>6If $`\delta =\pm \pi /2`$ or $`p_{ij}`$ is isotropic, $`p_{r\varphi }=0`$ and there is no transport. Since the denominator in equation (14) is positive, $$\text{sgn}p_{r\varphi }=\text{sgn}\delta .$$ (15) Outward (inward) transport corresponds to positive (negative) $`p_{r\varphi }`$, which in turn corresponds to positive (negative) $`\delta `$. In the principal axis basis, the viscous stress equations (1)–(4) read (cf. Goldreich & Tremaine 1978): $$_tp_{11}S\mathrm{\Omega }(\mathrm{sin}2\delta )p_{11}=\left(_tp_{11}\right)_c+\dot{H}_{11}$$ (16) $$_tp_{22}+S\mathrm{\Omega }(\mathrm{sin}2\delta )p_{22}=\left(_tp_{22}\right)_c+\dot{H}_{22}$$ (17) $$_tp_{33}=\left(_tp_{33}\right)_c+\dot{H}_{33}$$ (18) $$(p_{11}p_{22})_t\delta (2S\mathrm{sin}^2\delta )\mathrm{\Omega }p_{22}+(2S\mathrm{cos}^2\delta )\mathrm{\Omega }p_{11}=(p_{11}p_{22})(_t\delta )_c+\dot{H}_{12},$$ (19) where $`S=d\mathrm{ln}\mathrm{\Omega }/d\mathrm{ln}r`$. Equations (16) and (17) state that for $`S>0`$, the steady ($`_t=0`$) inward transport ($`\delta <0`$) of angular momentum demands that the net tendency of collisions and external stirring be to further lengthen the principal long axis of the velocity ellipsoid and to further shorten the principal short axis. Collisions between hard, imperfectly elastic spheres tend to isotropize the velocity ellipsoid: $`(_tp_{11})_c<0<(_tp_{22})_c`$. Thus, in the absence of stirring, transport is normally outwards in axisymmetric particle disks. ## 4 Angular Momentum Transport in Externally Stirred Disks In this section we solve for the steady state structure of a particle disk using a simple model for the effects of interparticle collisions and external stirring. We also discuss the fluid analogue of our particle disk model in some detail. ### 4.1 The Particle Scattering Function Following, e.g., Cook & Franklin (1964), Shu & Stewart (1985), and Narayan, Loeb, & Kumar (1994) we take the scattering function in the Boltzmann equation to be given by the Krook approximation: $$(_tf)_c=\frac{f_0f}{\tau },$$ (20) where $`\tau `$ is the collisional mean free time, $`f`$ is the particle distribution function, and $`f_0`$ is the “equilibrium” or “post-scattering” distribution function. The corresponding collision term in the viscous stress equations is $`\left(_tp_{ij}\right)_c=(p_{ij}^0p_{ij})/\tau `$. The physical idea behind the Krook model is that when particles collide they are removed from phase space at a rate $`f/\tau `$ and are returned to phase space with the new distribution function $`f_0`$. We assume that collisions do not generate off diagonal stresses so that $`p_{ij}^0=0`$ if $`ij`$. We account for the fact that particle collisions are inelastic, converting some fraction of the kinetic energy of the colliding particles into heat. With this model, the relevant second moments of $`f_0`$ are given by (see Kumar et al. 1995) $$p_{rr}^0=(1\xi _r)\frac{\sigma ^2}{3},p_{\varphi \varphi }^0=(1\xi _\varphi )\frac{\sigma ^2}{3},\mathrm{and}p_{zz}^0=(1\xi _z)\frac{\sigma ^2}{3},$$ (21) where $`\sigma `$ is the velocity dispersion ($`\sigma ^2=p_{rr}+p_{\varphi \varphi }+p_{zz}`$) and the parameters $`\xi _iϵ[0,1]`$ measure the inelasticity of the collisions; $`\xi 0`$ is elastic, while $`\xi 1`$ is highly inelastic. Equation (21) allows the inelasticity of colliding particles to be anisotropic, i.e., to be different for the $`r`$, $`\varphi `$, and $`z`$ directions. Such an anisotropic scattering function is unrealistic for real particle disks in nature (e.g., planetary rings); collisions between hard spheres at the microscopic level cannot be expected to respect the cylindrical coordinate system. Our strategy is to first solve for the structure of “real” externally stirred particle disks, in which the $`\xi _i`$’s are equal. This captures much of the physics in which we are interested. We will show, however, that the assumption of equal $`\xi _i`$’s does not furnish a fully adequate model of fluid disks and that the introduction of an “anisotropic inelasticity” can remedy this defect. ### 4.2 The Fluid Analogue of the Scattering Function Our particle scattering function has a mathematically rigorous analogue in fluid turbulence. Namely, it is equivalent to a particular closure scheme for the a priori unknown turbulent quantities $`ϵ_{ij}`$ and $`\pi _{ij}`$: $$ϵ_{ij}=\frac{T_{ij}}{\tau }\mathrm{and}\pi _{ij}=\frac{(1\xi _i)}{3\tau }\sigma ^2\delta _{ij}.$$ (22) Interpreted as such, $`\tau `$ is roughly the timescale on which free turbulence would decay (the eddy turnover time) and the inelasticity parameter, $`\xi _i`$, measures the efficiency with which fluctuating pressure forces perform work in the $`i`$th direction. The closure model proposed in equation (22) differs from conventional second-order closure models of the Reynolds stress equation (e.g., Speziale 1991; see Kato & Yoshizawa 1997 for an application to Keplerian disks). We will show, however, that this simple model provides considerable insight into the angular momentum transport properties of accretion disks. In contrast to $`\tau `$, it is not obvious what values to assign $`\xi _i`$ in hydrodynamic turbulence. Although in one sense turbulence is “inelastic,” with energy cascading to small scales in the absence of source terms, the physical meaning of $`\xi _i`$ defined by equation (22) is more precise: fluid turbulence is “inelastic” if and only if fluctuating pressure forces are inefficient at doing work. Throughout the remainder of this paper it is worth keeping in mind that numerical simulations suggest that the highly “inelastic” limit is appropriate for Keplerian disks. Note also that, unlike in the particle disk case, we would not expect the $`\xi _i`$’s to be identical in the fluid case. The fluctuating pressure forces are macroscopic in origin and thus can plausibly be different for the $`r`$, $`\varphi `$, and $`z`$ directions. ### 4.3 The Stress Tensor Equations With our model for the interparticle collision term, the steady state equations for the stress tensor become $$p_{rr}(1\xi _r)\frac{\sigma ^2}{3}4\mathrm{\Omega }\tau p_{r\varphi }=H_{rr},$$ (23) $$p_{\varphi \varphi }(1\xi _\varphi )\frac{\sigma ^2}{3}+\frac{\kappa ^2\tau }{\mathrm{\Omega }}p_{r\varphi }=H_{\varphi \varphi },$$ (24) $$p_{zz}(1\xi _z)\frac{\sigma ^2}{3}=H_{zz},$$ (25) and $$\frac{\kappa ^2\tau }{2\mathrm{\Omega }}p_{rr}2\mathrm{\Omega }\tau p_{\varphi \varphi }+p_{r\varphi }=0,$$ (26) where $`H_{ij}\tau \dot{H}_{ij}`$, i.e., the net stirring over the mean free time. An “energy equation” can be derived by summing equations (23)–(25): $$HH_{rr}+H_{\varphi \varphi }+H_{zz}=\xi \sigma ^2+2p_{r\varphi }\mathrm{\Omega }^{}\tau ,$$ (27) where $`\mathrm{\Omega }^{}=d\mathrm{\Omega }/d\mathrm{ln}r`$ and $`3\xi =\xi _r+\xi _\varphi +\xi _z`$. For $`\mathrm{\Omega }^{}<0`$, equation (27) shows that, if there is no external stirring, transport must be outwards or zero in steady state. This is because any energy lost due to inelasticity must be extracted from the shear in the mean flow, which requires $`p_{r\varphi }>0`$. No such thermodynamic constraint on the direction of angular momentum transport exists if there is external stirring. ### 4.4 Transport in a Keplerian Disk The general solution to equations (23)–(26) is given in the Appendix. Here we focus on the case of a Keplerian disk, for which $`\kappa =\mathrm{\Omega }`$; in the next subsection we briefly contrast the Keplerian disk with a constant angular momentum disk. We define $`x=\mathrm{\Omega }\tau `$ and solve for the $`r\varphi `$ component of the stress tensor: $$p_{r\varphi }=\frac{x\sigma ^2}{2}\left(\frac{\frac{4}{3}\left[1\xi _\varphi \right]\frac{1}{3}\left[1\xi _r\right]\stackrel{~}{H}_{rr}\xi +4\stackrel{~}{H}_{\varphi \varphi }\xi }{1+x^2\left[2.5\stackrel{~}{H}_{rr}+10\stackrel{~}{H}_{\varphi \varphi }+4\stackrel{~}{H}_{zz}\right]}\right),$$ (28) where $`\stackrel{~}{H}_{ii}=H_{ii}/H`$. The velocity dispersion of the disk is not an independent parameter but is determined by solving the equation $$\sigma ^2=\frac{H+x^2\left(2.5H_{rr}+10H_{\varphi \varphi }+4H_{zz}\right)}{(1+4x^2)\xi 0.5\xi _rx^2+2\xi _\varphi x^21.5x^2}.$$ (29) #### 4.4.1 Real Particle Disks: Isotropic Inelasticity For a real particle disk with isotropic inelasticity ($`\xi _r=\xi _\varphi =\xi _z=\xi `$) equations (28) and (29) become $$p_{r\varphi }=\frac{x\sigma ^2}{2}\left(\frac{1\xi \stackrel{~}{H}_{rr}\xi +4\stackrel{~}{H}_{\varphi \varphi }\xi }{1+x^2\left[2.5\stackrel{~}{H}_{rr}+10\stackrel{~}{H}_{\varphi \varphi }+4\stackrel{~}{H}_{zz}\right]}\right)$$ (30) and $$\sigma ^2=\frac{H+x^2\left(2.5H_{rr}+10H_{\varphi \varphi }+4H_{zz}\right)}{(1+5.5x^2)\xi 1.5x^2}.$$ (31) In the limit of no external stirring, we recover the usual results of particle disk theory. In such an isolated disk, the inelasticity adjusts to the optical depth of the disk, $$\xi =\frac{1.5x^2}{1+5.5x^2}.$$ (32) This follows from equation (31) by setting $`H=0`$ and requiring finite $`\sigma ^2`$. Since the collisional inelasticity depends on the relative impact velocity, equation (32) determines a velocity dispersion (Goldreich & Tremaine 1978); call this $`\sigma _0`$. Inserting equation (32) for $`\xi `$ in equation (30), we confirm that the transport is always outwards, with<sup>7</sup><sup>7</sup>7To derive eq. , write eq. as $`p_{r\varphi }=0.5x\sigma ^2N/D`$. Rewritting the denominator as $`D=1+4x^2+x^2(6\stackrel{~}{H}_{\varphi \varphi }1.5\stackrel{~}{H}_{rr})`$ and substituting eq. for $`\xi `$ into the numerator, $`N`$, all terms $`H_{ij}`$ cancel, yielding eq. . $$p_{r\varphi }=\frac{\xi \sigma ^2}{3x}=\frac{x\sigma ^2}{2+11x^2}.$$ (33) The fact that isolated fluid disks appear to be hydrodynamically stable can be explained mechanistically in the sense that they do not satisfy equation (32). For any $`x`$, equation (32) has a solution only if $`\xi ϵ[0,0.27]`$. Simulations suggest that the work done by fluctuating pressure forces is far less efficient than this. Therefore, for $`H=0`$, the only solution to equation (31) is $`\sigma _0^2=0`$, i.e., no turbulence. What happens if the disk is externally stirred? For $`H\xi \sigma _0^2`$ the disk structure is only weakly affected by the stirring; corrections to $`p_{r\varphi }`$, e.g., are $`O[H/\xi \sigma _0^2]1`$. Physically, this is because, for $`\sigma _00`$, heating must become significant relative to the “natural” velocity dispersion in order to modify the disk structure. For a hydrodynamically stable fluid disk with $`\sigma _0=0`$, however, the “natural” velocity dispersion vanishes; thus, for any $`H`$, the properties of the external stirring (together with the “collisions”) determine the angular momentum transport properties of the disk. In the presence of significant stirring, i.e., H >σ02 >𝐻subscriptsuperscript𝜎20H\mathrel{\mathchoice{\vbox{\offinterlineskip\halign{$\m@th\displaystyle\hfil#\hfil$\cr>\crcr\sim\crcr} }}{\vbox{\offinterlineskip\halign{$\m@th\textstyle\hfil#\hfil$\cr>\crcr\sim\crcr} }}{\vbox{\offinterlineskip\halign{$\m@th\scriptstyle\hfil#\hfil$\cr>\crcr\sim\crcr} }}{\vbox{\offinterlineskip\halign{$\m@th\scriptscriptstyle\hfil#\hfil$\cr>\crcr\sim\crcr} }}}\sigma^{2}_{0}, the direction of angular momentum transport can reverse; in particular, the sign of the angular momentum transport is then determined by the sign of the numerator in equation (30). Note that the sign of $`p_{r\varphi }`$ does not explicitly depend on $`x=\tau \mathrm{\Omega }`$. This is consistent with the arguments given in §2 and §3; only the properties of the collisions and the stirring determine the direction of angular momentum transport, not the degree to which the disk is collisional. From the numerator in equation (30) it follows that (1) energy input along $`r`$ tends to lead to inward transport ($`H_{rr}`$ enters with a negative sign), (2) energy input along $`\varphi `$ tends to lead to outward transport ($`H_{\varphi \varphi }`$ enters with a positive sign), and (3) elastic scattering tends to lead to outward transport ($`p_{r\varphi }>0`$ if $`\xi 1`$). All of these results are readily understood using the theorem of §3. The principal long axis of the velocity ellipsoid generally has a much larger projection along the radial direction than the azimuthal direction.<sup>8</sup><sup>8</sup>8In the collisionless unstirred limit, equation (19) states that $`p_{rr}/p_{\varphi \varphi }=2/(2S)=4`$ for steady-state Keplerian disks. Thus $`r`$ stirring corresponds to elongating the long axis of the velocity ellipsoid and so should (and does) lead to inward transport (and vice-versa for $`\varphi `$ stirring). The transport only respects the direction of external stirring, however, if the collisions are fairly inelastic. In the elastic limit, collisions efficiently transfer energy from one direction to another, isotropizing the distribution function; transport is then always outwards. Two special cases of equation (30) are of particular interest: 1. $`R`$ stirring only: This is relevant for interpreting the work of Stone, Pringle, & Begelman (1999) and Igumenshchev & Abramowicz (1999) who carried out simulations of advection-dominated accretion flows, which are radially convective. In these flows, one can treat the convection as “externally” driven by the viscous dissipation associated with magnetic fields. As discussed by Narayan, Igumenshchev, & Abramowicz (2000) and Quataert & Gruzinov (2000), the convective angular momentum transport is inwards in these simulations, with $`\alpha `$ as large as $`0.1`$ ($`\alpha p_{r\varphi }/\sigma ^2`$ is the dimensionless viscosity). As noted above, it is particularly natural for radial convection to produce inward transport since radial convection tends to directly expand the long axis of the velocity ellipsoid. More concretely, equation (30) predicts that the transport should be inwards in the presence of $`r`$ stirring so long as $`\xi >1/2`$. Moreover, for $`\xi 1`$ and $`x1`$,<sup>9</sup><sup>9</sup>9This means that the eddy turnover time is comparable to the rotational period of the disk, as is indeed the case. $`\alpha `$ should be $`0.1`$ in the presence of $`r`$ stirring, consistent with simulations. 2. $`Z`$ stirring only: This is relevant for interpreting the simulations of Stone & Balbus (1996) in which vertical convection in an otherwise stable Keplerian disk generated low levels of inward angular momentum transport ($`\alpha 10^4`$). For a particle disk, equation (30) shows that transport is always outwards in the presence of $`z`$ stirring: $`p_{r\varphi }(1\xi )`$. This can be understood by noting that the $`z`$ dynamics in thin reflection-symmetric disks is largely decoupled from the $`r`$ and $`\varphi `$ dynamics which determine the direction of transport. In particular, energy input in the $`z`$ direction is transferred to the $`r`$ and $`\varphi `$ directions by particle collisions, which attempt to isotropize the distribution function. By the theorem of §3 the transport must then be outwards. #### 4.4.2 “Anisotropic Inelasticity”: Inward Transport with Vertical Stirring The analysis of the previous section applies to real particle disks in which the inelasticity is isotropic. Interestingly, it shows that vertically stirred particle disks would transport angular momentum outwards, opposite to what is found in simulations of hydrodynamic turbulence. As noted in §4.2, the fluid analogue of inelasticity is inefficient fluctuating pressure forces; such forces are macroscopic in origin and therefore need not be the same in the $`r`$, $`\varphi `$, and $`z`$ directions. We now discuss how such “anisotropic inelasticity” can produce inward transport in the presence of vertical stirring. In the presence of vertical stirring only, equation (28) reduces to $$p_{r\varphi }=\frac{x\sigma ^2}{2}\left(\frac{\frac{4}{3}\left[1\xi _\varphi \right]\frac{1}{3}\left[1\xi _r\right]}{1+4x^2}\right).$$ (34) From the numerator in equation (34) it follows that elasticity in the $`r`$ direction tends to lead to inward transport ($`1\xi _r`$ has a negative prefactor), while elasticity in the $`\varphi `$ direction tends to lead to outward transport ($`1\xi _\varphi `$ has a positive prefactor). Inward transport in the presence of $`z`$ stirring requires $`\xi _\varphi >0.75+0.25\xi _r`$. Physically, this means that fluctuating pressure forces are particularly inefficient at doing work in the $`\varphi `$ direction (cf. Stone & Balbus 1996). In fact, in the absence of explicit $`\varphi `$ stirring, equation (28) shows that the transport is always inwards if $`\xi _\varphi 1`$.<sup>10</sup><sup>10</sup>10This is analogous to Stone & Balbus’s argument that axisymmetric simulations of hydrodynamic turbulence will always find inward transport. In order to reproduce $`\alpha 10^4`$ in the presence of $`z`$ stirring, as found in the simulations, equation (34) requires (taking $`x1`$) $`1\xi _r10^3`$, and $`1\xi _\varphi 10^3`$. ### 4.5 Transport in a Constant Angular Momentum Disk Balbus and collaborators have emphasized the interesting point that a constant angular momentum, $`\kappa =0`$, disk is analogous to a linear shear flow (e.g., Balbus, Hawley, & Stone 1996). Indeed, numerical simulations show that the $`\kappa =0`$ disk is unstable to nonlinear perturbations and produces outwards angular momentum transport. This feature is also readily captured by our model. By equation (A1) the $`r\varphi `$ stress in a $`\kappa =0`$ disk is given by $$p_{r\varphi }=2x\sigma ^2\left(\frac{\frac{1\xi _\varphi }{3}+\xi \stackrel{~}{H}_{\varphi \varphi }}{1+8x^2\stackrel{~}{H}_{\varphi \varphi }}\right),$$ (35) which is positive definite. For unstirred $`\kappa =0`$ disks in which the particle inelasticity is isotropic, $$\xi =\frac{8x^2}{3+8x^2}.$$ (36) In contrast to equation (32) for a Keplerian disk, equation (36) can be satisfied for some $`x`$ for any $`\xi ϵ[0,1]`$. For a fluid disk this means that, regardless of the efficiency of fluctuating pressure forces, outward transport can be sustained in an unstirred disk. This is consistent with the fact that shear flows are nonlinearly unstable. ## 5 Discussion This paper explores the angular momentum transport properties of axisymmetric, externally stirred particle disks as a means of better understanding hydrodynamic turbulence in fluid disks. We are not, however, advocating that turbulence is a sort of macroscopic kinetic theory, with colliding turbulent “blobs” literally replacing colliding particles. Rather, we argue that it is physically revealing to exploit the similar mathematics governing turbulent fluid disks and particle disks. There are one-to-one correspondences between terms in the Boltzmann stress tensor equation governing particle disks and terms in the Reynolds stress tensor equation governing fluid disks (§2). This analogy reveals that the role of differential rotation and epicyclic motion is the same in both fluid and particle disks; both the rotation rate and the epicyclic frequency couple to the components of the stress tensor identically in the two cases. Consequently, the transport properties of fluid disks can differ from those of particle disks only to the extent that the “scattering function” for fluid elements differs from that of particles. The formal, and very useful, analogy is that interparticle collisions in particle disks play the same role as viscous dissipation and fluctuating pressure forces in turbulent disks. Both transfer energy in random motions associated with one direction to those associated with another, and convert kinetic energy into heat. Motivated by the above considerations, we solve for the structure of a particle disk using a simple model for the collision term in the Boltzmann equation (§4). We allow different collisional inelasticities along the $`r`$, $`\varphi `$, and $`z`$ directions. Although an anisotropic inelasticity is not realistic for actual particle collisions, its introduction serves to make the analogy with fluid turbulence complete. The fluid analogue of particle inelasticity is the inefficacy with which fluctuating pressure forces do work, and there is no a priori reason for the latter to be isotropic. A final ingredient of our model is the insertion of an external stirring mechanism, modeled as constant source terms in the $`r`$, $`\varphi `$, and $`z`$ energy equations. External stirring is relevant for interpreting numerical simulations in which either vertical or radial convection is present (e.g., Stone & Balbus 1996; Stone, Pringle, & Begelman 1999). We believe that this particle model captures the transport properties of fluid disks. With a single equation for the stress tensor (eq. ) we can reproduce results from standard particle disk theory and from a number of numerical simulations of hydrodynamic turbulence. In the limit of no external stirring, differential rotation is the only source of free energy in our disks. The angular momentum transport induced by turbulence or particle collisions must then be outwards or zero in steady state. This is because both particle collisions and turbulence generate heat, which can only be compensated for by a positive $`r\varphi `$ stress (see eq. ). To have outward transport in a Keplerian disk requires that “collisions” be relatively efficient at isotropizing the distribution function (see eq. and the associated discussion). This tendency towards isotropization is normally satisfied in axisymmetric particle disks such as planetary rings; these systems therefore transport momentum outwards while maintaining some degree of random motion (turbulence). By contrast, the hydrodynamic stability of isolated fluid disks can be understood by noting that, according to simulations, fluctuating pressure forces are extremely inefficient at isotropization. Thus there can be no self-sustaining turbulence in these systems.<sup>11</sup><sup>11</sup>11Kato & Yoshizawa (1997) have argued that if pressure fluctuations in turbulent Keplerian disks behave like they do in shear flows and other turbulent laboratory fluids, the isotropization is sufficient to yield self-sustaining turbulence. This assumes, however, that Reynold’s stress closure models developed for laboratory problems are applicable to Keplerian disks. In the presence of significant external stirring, there are no thermodynamic constraints on the direction of angular momentum transport. Instead, the sign of the transport is determined by the nature of the stirring and the properties of the collision term in the Boltzmann equation (or its suitably defined analogue in the fluid problem); formally, a necessary and sufficient condition for inward transport is that the combined tendency of collisions and stirring be to make the long equatorial axis of the velocity ellipsoid yet longer and the short equatorial axis yet shorter (§3). More physically, for any mechanism of external stirring (e.g., convection, spiral shocks, etc.), our model predicts the direction of transport as a function of the “inelasticity” of collisions (eq. ). The latter is, of course, not known a priori for the fluid problem, which underscores the parametrized approach of our model. Nonetheless, it is gratifying that, so long as collisions are suitably inelastic, our model finds inward transport for both vertical and radial stirring (e.g., convection), with the transport naturally much more efficient for the latter than for the former. Inward transport in the presence of purely vertical stirring requires, however, an anistropic “inelasticity,” for which “collisions” are more inelastic in the azimuthal direction than in the radial direction (§4.4.2). By equating inelasticity with inefficient fluctuating pressure forces, as advocated above, we find results that are consistent with analytic arguments and simulations of hydrodynamic turbulence (e.g., Balbus & Hawley 1998). To conclude, we note that our analysis expands on several aspects of the Boltzmann equation model for radial convection in rotating media considered by Kumar, Narayan, & Loeb (1995). These authors treated “collisions” between convective eddies using the same scattering function we have considered (eq. ) and found that for sufficiently inelastic collisions, radial convection would produce inward angular momentum transport. Interestingly, we now see that this inward transport has little to do with the details of convectively unstable fluids (e.g., the superadiabatic entropy gradient, correlated temperature and velocity fluctuations, etc. analyzed by Kumar et al.). Rather, it is a generic property of disks in the presence of external “stirring” of the radial velocity dispersion. Moreover, we have emphasized the analogy between inelastic collisions in particle disks and inefficient fluctuating pressure forces in turbulent disks, better establishing the relevance of kinetic theory models to the fluid problem. ###### Acknowledgements. We thank Steve Balbus, Roger Blandford, Peter Goldreich, Jeremy Goodman, Pawan Kumar, Ramesh Narayan, Scott Tremaine, and an anonymous referee for useful discussions and comments. EQ is supported by NASA through Chandra Fellowship PF9-10008, awarded by the Chandra X–ray Center, which is operated by the Smithsonian Astrophysical Observatory for NASA under contract NAS 8-39073. EC acknowledges partial support from a Caltech Kingsley Foundation Fellowship. ## Appendix A Solution for General $`\kappa `$ Defining $`x=\mathrm{\Omega }\tau `$, $`y=\kappa \tau `$, and $`x^{}=dx/d\mathrm{ln}r=(y^24x^2)/2x`$, one can solve equations (23)–(26) to find: $$p_{r\varphi }=x\sigma ^2\left(\frac{\frac{2}{3}\left[1\xi _\varphi \right]\frac{y^2}{6x^2}\left[1\xi _r\right]\stackrel{~}{H}_{rr}\frac{\xi y^2}{2x^2}+2\stackrel{~}{H}_{\varphi \varphi }\xi }{1+4y^2+\frac{y^2x^{}}{x}\stackrel{~}{H}_{rr}4xx^{}\stackrel{~}{H}_{\varphi \varphi }}\right)$$ (A1) and $$\sigma ^2=\frac{H+4Hy^2+\frac{y^2x^{}}{x}H_{rr}4xx^{}H_{\varphi \varphi }}{(1+4y^2)\xi \frac{x^{}}{3}\left[\frac{y^2}{x}\left(1\xi _r\right)4x\left(1\xi _\varphi \right)\right]}.$$ (A2)
warning/0003/astro-ph0003371.html
ar5iv
text
# On the formation and evolution of the globular cluster 𝜔 Centauri ## 1 Introduction $`\omega `$ Centauri is one of the most interesting globular cluster in the Milky Way, and maybe the most studied one (Majewski et al. 1999, Lee et al. 1999, Pancino 1998). A compilation of its main properties is given in Table 1. The most striking feature of this cluster is the measured metallicity spread, which has been interpreted as the evidence of a multiple stellar population inside it (Norris et. al 1996, Suntzeff & Kraft 1996). The precise metallicity distribution function (MDF) is nonetheless still disputed. Norris et al (1996) suggest the presence of a secondary peak at $`[Ca/H]0.9`$ roughly 5 times smaller that the main peak at $`[Ca/H]1.4`$. This trend is not confirmed by Suntzeff & Kraft (1996), who claim for a more regular MDF. On the base of a large photometric survey Lee et al. (1999) (but see also Ortolani et al 1999) analyzed the color distribution of a sample of bright stars, showing that on the average it has an e-folding trend, with the presence of several significant metallicity peaks. However Majewski et al. (1999) arrive at a somewhat different conclusion, showing that the MDF has a gaussian shape with the maximum at $`[Fe/H]1.7`$, and with some evidences of a secondary peak. Although different, all these analyses point to the common picture of an object which experienced an irregular self-enrichment over its evolution. Putting together the chemical and kinematical properties Majewski et al. (1999) claim that $`\omega `$ Centauri might be a possible dwarf galaxy relict. In this paper we propose a N-body/gasdynamical model for the formation and evolution of $`\omega `$ Centauri, suggesting that this globular cluster can actually be the remnant of a dwarf elliptical galaxy, formed and evolved avoiding strong mergers, and eventually captured by the Milky Way. To this aim, the plan of this paper is as follows. In Sect. 2 we describe our model and the initial conditions setup; Sect. 3 to 5 are dedicated to the analysis of the structure, chemistry and internal kinematics, respectively, of our model, and the comparison with $`\omega `$ Centauri; in Sect. 6 we investigate about the possible presence of an extended DM halo around the cluster. Finally Sect. 7 summarizes our results. ## 2 Technique and Initial Conditions The simulation presented here has been performed using the Tree–SPH code developed by Carraro at al. (1998) and Buonomo et al. (2000). In this code, the properties of the gas component are described by means of the Smoothed Particle Hydrodynamics (SPH) technique, whereas the gravitational forces are computed by means of a hierarchical tree algorithm using a tolerance parameter $`\theta =0.8`$ and expanding tree nodes to quadrupole order. We adopt a Plummer softening parameter. In SPH each particle represents a fluid element whose position, velocity, energy, density etc. are followed in time and space. The properties of the fluid are locally estimated by an interpolation which involves the smoothing length $`h_i`$. In our code each particle possesses its own time and space variable smoothing length $`h_i`$, and evolves with its own time-step. This renders the code highly adaptive and flexible, and suited to speed-up the numerical calculations. Radiative cooling is described by means of numerical tabulations as a function of temperature and metallicity taken from Sutherland & Dopita (1993) and Hollenbach & McKee (1979). This allows us to account for the effects of variations in the metallicity among the fluid elements and for each of these as a function of time and position. The chemical enrichment of the gas-particles caused by Star Formation (SF) and stellar ejecta is described by means of the closed-box model applied to each gas-particle (cf. Carraro et al. 1998 for more details). SF and Feed-back algorithm are described in great detail by Buonomo et al. (2000). In brief, SF is let depend on the total mass density - baryonic (gas and stars) and dark matter \- of the system and on the metal–dependent cooling efficiency. Moreover we consider the effects of energy (and mass) feed-back from SNe of type II and Ia, and stellar winds from massive stars. Instead of selecting halos from large cosmological N-body simulations, we opted to set up a protogalaxy as an isolated virialized DM halo, whose density profile is: $$\rho (r)\frac{1}{r}.$$ DM particles are distributed inside the sphere according to an acceptance–rejection procedure. Along with positions, we assign also velocities according to: $$v(r)\sqrt{(}r\times ln(\frac{1}{r})).$$ which is the solution of the Jeans equation for spherical isotropic collisionless system for the adopted density profile. The system is let evolve until virial equilibrium is reached. Gas particles are then distributed homogeneously inside the DM halo with zero velocity field, thus mimicking the cosmological gas infall inside the DM potential well (White & Rees 1978). For the simulation presented here we consider a $`9\times 10^7M_{}`$ DM halo, sampled by 10,000 particles, homogeneously filled with $`1\times 10^7M_{}`$ baryonic mass in form of metal poor ($`Z=10^4`$) gas, sampled by 10,000 particles. This way each gas particle has a mass of $`10^3M_{}`$. ## 3 The structure The evolution of our model is shown in Fig. 1, where we plot the SF as a function of time. Starting from a $`10^4`$ <sup>o</sup>K metal poor gas, cooling drives baryons towards the center of the DM potential well. Part of this gas (about 20%) is transformed into stars with a very irregular SF history, whereas the remaining 80% is thrown away via a galactic wind mechanism, for which SNe and stellar winds are responsible. The final stars distribution resembles a triaxial object, with $`a/b=0.859`$ and $`a/c=0.967`$. Correspondingly the ellipticity $`(ϵ=1b/a)`$ turns out to be 0.141, not far from the mean value reported by Geyer et al. (1983). The stars distribution is shown in Fig. 2 (onto the X-Z plane) and 3 (onto the X-Y plane). For the sake of an easier comparison with $`\omega `$ Centauri the model has been placed at the cluster actual distance (5.3 kpc). The total mass in stars amount to about $`2\times 10^6M_{}`$, sampled by about 2,000 star particles. The spherically averaged mass density profile derived from our simulation is shown in Fig. 4 . Each filled circle represents the stars density $`\rho _i`$ computed as the mass inside a spherical shell $`i`$ with size the softening $`ϵ`$, divided by the shell volume. For each density estimate $`\rho _i`$ we plot as error bar the uncertainty of the density estimate evaluated as the poissonian error related to the shell population. We perform a fit with a King (1962) profile to check whether our model recovers the structural properties of $`\omega `$ Centauri. The cluster indeed (see Table 1) has a core and tidal radius $`r_c`$ and $`r_t`$ equal to 2.6 and 45 arcmin, respectively. By using a King (1962) model (solid overimposed line), we found the best fit for $`r_c`$= 3.2 and $`r_t`$= 65 arcmin, which implies a concentration $`c=log(\frac{r_t}{r_c})`$ = 1.30, somewhat higher than the value reported in Table 1. While the value of $`r_c`$ we obtain agrees with the observed one, the value of $`r_t`$ turns out to be much larger than the measured one for $`\omega `$ Centauri. In other words we over-estimate the tidal radius. In our opinion this is due to the fact that in our simplified model we neglect the effect of the tidal stripping that the cluster probably did experience when it was accreted by the Milky Way. This way the outermost stars were probably removed reducing the tidal radius. Finally, the central density, computed assuming the $`r_c`$ derived above, is about 2660 $`M_{}/pc^3`$. ## 4 Internal kinematics The internal kinematics of $`\omega `$ Centauri has been analyzed in great detail by Merritt et al. (1997) using the radial velocity catalogue obtained with CORAVEL by Mayor et al. (1997). From this study it emerges that the cluster has a peak rotational speed of about 8 km/sec at 11 pc from the center. Moreover the cluster has a central velocity dispersion $`\sigma `$ = 17 km/sec, the higher one among globular clusters (see Fig. 8 in Merritt et al 1997). From the analysis of our data we find evidences of smaller rotational speed (about 4 km/sec at 15 pc), whereas the velocity dispersion profile $`\sigma (r)`$ (see Fig. 5) shows a central value of about 13 km/sec, denoting that our model is somewhat colder than $`\omega `$ Centauri. ## 5 Metallicity distribution The stars MDF at the end of the simulation is presented in Fig. 6. The bulk of stars has a metallicity in the range between $`\text{[Fe/H]}=1.9`$ and $`\text{[Fe/H]}1.6`$. A secondary peak is observed at $`\text{[Fe/H]}=1.2`$, about 4 times smaller than the major peak. Finally a significant number of stars has a relatively high metallicity around $`\text{[Fe/H]}=0.50`$. The global shape resembles the e-folding SF history shown in Fig. 1, with secondary peaks which mark successive bursts of star formation. The good agreement we find with the data presented by Majewski et al. (1999) confirms their suggestion that $`\omega `$ Centauri experienced an irregular self-enrichment over its evolution and may actually be the core of a larger dwarf elliptical galaxy. ## 6 Dark Matter around $`\omega `$ Centauri? Our simulation starts with a virialised DM halo, in whose center gas concentrates due to cooling instability, forming stars. The final stars and DM distribution is shown in Fig. 7, where open squares refer to DM, and filled circles represent stars. The inner region of the cluster is dominated by stars up to 20 arcmin, which we call the transition radius $`r_{tr}`$. In the outer region DM dominates over a distance 6 times larger than the transition radius. We expect that the internal kinematics of stars in the inner region is unlikely to be influenced by DM. On the contrary, stars outside the transition radius $`r_{tr}`$ (20 arcmin) should show a kinematics strongly influenced by DM. A detail spectroscopic study of the stars kinematics in the cluster envelope should be able to confirm or deny the presence of DM. There is indeed in our model a trend of the velocity dispersion to weakly increase out of the transition radius. ## 7 Conclusions We have presented a N-body/gasdynamical simulation of the formation and evolution of the globular cluster $`\omega `$ Centauri. We are able to reproduce the bulk properties of the cluster, namely structure, kinematics and chemistry assuming that it formed and evolved in isolation, and eventually fell inside the Milky Way potential well. According to our results and to the dwarf galaxies taxonomy proposed by Roukema (1999), $`\omega `$ Centauri can actually be a cosmological dwarf by mass, formed in a high redshift low mass halo, which escaped significant merging up to the present time. Finally we stress that in order to obtain these results, an extended DM halo should surround the present day $`\omega `$ Centauri. ### Acknowledgments G.C. expresses his gratitude to Prof. S. Ortolani for fruitful discussions. The referee, George Meylan, is acknowledged for important suggestions which led to the improvement of the paper. This work has been financed by italian MURST and ASI. ## References Buonomo F., Carraro G., Chiosi C., Lia C., 2000, MNRAS 312, 371 Carraro G., Lia C., Chiosi C., 1998, MNRAS 291, 1021 Geyer E.H. Hopp U., Nelles B., 1983, A&A 125, 359 Harris W. E. 1996, AJ 112, 1487 Hollenbach D., McKee C.F., 1979, ApJS 41, 555 King I., 1962, AJ 67, 471 Lee Y.-W., Joo J.-M., Sohn Y.-J., Lee H.-C., Walker A.R., 1999, Nat., in press (astro-ph/9911137) Majewski S.R., Patterson R.J., Dinescu D.I. et al., 1999, in ”The Galactic Halo: from Globular Clusters to Field Stars”, 35th Liege Astrophysical Colloquium (astro-ph/9910278) Mayor M., Meylan G., Udry S. et al., 1997, AJ 114, 1087 Merritt D., Meylan G., Mayor M., 1997, AJ 114, 1074 Ortolani S., Covino S., Carraro G., 1999, in ”The Third Stromlo Simposium: The Galactic Halo”, Eds Gibson B.K., Axelrod T.S., Putman M.E., ASP Conference Series Vol. 165, 316 Norris J.E., Freeman K.C., Mighell K.J., 1996, ApJ 462, 251 Pancino E., 1998, Master Thesis, Padova University Roukema B.F., 1999, in ”XVI Rencontre de Moriond: Dwarf Galaxies and Cosmology”, in press (astro-ph/9806111) Searle L., Zinn R., 1978, ApJ 225, 357 Suntzeff N.B., Kraft R.P., 1996, AJ 111, 1913 Sutherland R.S., Dopita M.A., 1993, ApJS 88, 253 White S.D.M., Rees M., 1978, MNRAS 183, 341
warning/0003/cond-mat0003274.html
ar5iv
text
# Spin-Gap States of a Periodic Mixed Spin Chain ## 1 Introduction Since the discovery of cuprate superconductors, quantum spin systems have been studied to understand the magnetic properties of their undoped mother materials. In particular quantum spin systems with spin gap are interesting since the spin gap is possibly related to the superconducting gap in the doped case. Although the materials are two-dimensional, it is basically important to study magnetic properties of spin gap systems in any dimensions. In one dimension, Haldane predicted that the spin excitation spectrum of a uniform spin chain is gapless if the magnitude of a spin is a half-odd-integer, and is gapful if it is an integer. The prediction has been confirmed theoretically and experimentally in many standpoints. Since Haldane’s prediction is based on a mapping of a spin Hamiltonian to the nonlinear $`\sigma `$ model (NLSM), it is recognized that the NLSM is a useful tool to investigate spin systems. The NLSM has been developed for periodic inhomogeneous spin chains. Affleck derived and examined an NLSM for a spin chain with bond alternation. The application of the NLSM to spin chains with more than one spin species in the period of more than two lattice spacings are also considered . In particular we have succeeded to derive an NLSM, with accurately keeping the degrees of freedom of the spin variables, for a general mixed spin chain with arbitrary finite period . We applied the NLSM method to systems with period 4 and spin species 2 (we defines the magnitudes as $`s_a`$ and $`s_b`$) . Imposing the condition that the ground state is singlet, only the possible order of spin magnitudes in a unit cell is of $`s_a`$-$`s_a`$-$`s_b`$-$`s_b`$. We obtained phase diagrams in the parameter space of the exchange couplings. In a few special cases, numerical calculations have been performed . The phase diagrams obtained by the NLSM are qualitatively agree with the numerical results. In this paper we examine another mixed spin chain with period 4 which includes three species of spins. The sequential order of magnitudes of spins in a unit cell is 1/2, 1, 3/2 and 1. There are four kinds of nearest-neighbour exchange couplings among them. If we consider a simplified case, the exchange parameters on the both sides of a 1/2 spin are the same value $`J`$, and those on the both sides of a 3/2 spin are of the same value $`J^{}`$. We in fact consider the distortion of $`J^{}`$ on the both sides of a 3/2 spin. We map the spin Hamiltonian describing the spin chain to an NLSM following Ref. . From the value of the topological angle in the NLSM, we determine the phase diagram of this model in the space of the exchange parameters. ## 2 Mapping to the nonlinear $`\sigma `$ model The Hamiltonian for the present spin chain is written as $`H`$ $`=`$ $`{\displaystyle \underset{j}{}}(J_1𝐒_{4j+1}𝐒_{4j+2}+J_2𝐒_{4j+2}𝐒_{4j+3}`$ (1) $`+`$ $`J_3𝐒_{4j+3}𝐒_{4j+4}+J_4𝐒_{4j+4}𝐒_{4j+5}),`$ where the magnitudes of $`𝐒_{4j+1}`$, $`𝐒_{4j+2}`$, $`𝐒_{4j+3}`$ and $`𝐒_{4j+4}`$ are $`(s_1,s_2,s_3,s_4)=({\displaystyle \frac{1}{2}},1,{\displaystyle \frac{3}{2}},1).`$ (2) A unit cell is illustrated in Fig. 1. To reduce the number of parameters, we restrict and reparametrize the antiferromagnetic exchange parameters as $`J_1=J_4=J,J_2={\displaystyle \frac{J^{}}{1\gamma }},J_3={\displaystyle \frac{J^{}}{1+\gamma }},`$ (3) where $`\gamma `$ ($`1<\gamma <1`$) is the distortion parameter describing the asymmetry between the couplings of the both sides of a spin with magnitude 3/2. Following Ref. , we can map a general periodic mixed spin chain satisfying a restriction to the NLSM. The mapped NLSM action is given by $`S_{\mathrm{eff}}=`$ $`{\displaystyle }d\tau {\displaystyle }dx\{i{\displaystyle \frac{J^{(0)}}{J^{(1)}}}𝐦({\displaystyle \frac{𝐦}{\tau }}\times {\displaystyle \frac{𝐦}{x}})`$ (4) $`+{\displaystyle \frac{1}{2aJ^{(1)}}}\left({\displaystyle \frac{J^{(1)}}{J^{(2)}}}{\displaystyle \frac{J^{(0)}}{J^{(1)}}}\right)\left({\displaystyle \frac{𝐦}{\tau }}\right)^2`$ $`+{\displaystyle \frac{a}{2}}J^{(0)}\left({\displaystyle \frac{𝐦}{x}}\right)^2\},`$ where $`𝐦`$ is an O(3) unit-vector field and $`a`$ is the lattice constant. For a general spin chain with period $`2b`$, the constants in the action are given as follows: $`{\displaystyle \frac{1}{J^{(n)}}}`$ $`=`$ $`{\displaystyle \frac{1}{2b}}{\displaystyle \underset{q=1}{\overset{2b}{}}}{\displaystyle \frac{(\stackrel{~}{s}_q)^n}{J_qs_qs_{q+1}}}(n=0,1,2)`$ (5) with accumulated spins $`\stackrel{~}{s}_q`$ $`=`$ $`{\displaystyle \underset{k=1}{\overset{q}{}}}(1)^{k+1}s_k.(q=1,2,\mathrm{},2b)`$ (6) The action (4) is of the standard form of the NLSM: $`S_{\mathrm{st}}`$ $`=`$ $`{\displaystyle }d\tau {\displaystyle }dx\{i{\displaystyle \frac{\theta }{4\pi }}𝐦({\displaystyle \frac{𝐦}{\tau }}\times {\displaystyle \frac{𝐦}{x}})`$ (7) $`+{\displaystyle \frac{1}{2gv}}\left({\displaystyle \frac{𝐦}{\tau }}\right)^2+{\displaystyle \frac{v}{2g}}\left({\displaystyle \frac{𝐦}{x}}\right)^2\},`$ where $`\theta `$ is the topological angle, $`g`$ is the coupling constant and $`v`$ is the spin wave velocity. In the present model, the accumulated spins (6) are $`(\stackrel{~}{s}_1,\stackrel{~}{s}_2,\stackrel{~}{s}_3,\stackrel{~}{s}_4)=({\displaystyle \frac{1}{2}},{\displaystyle \frac{1}{2}},1,0)`$ (8) and then the constants (5) are calculated as $`{\displaystyle \frac{1}{J^{(0)}}}={\displaystyle \frac{1}{J}}+{\displaystyle \frac{1}{3J^{}}},`$ $`{\displaystyle \frac{1}{J^{(1)}}}={\displaystyle \frac{1}{4}}\left({\displaystyle \frac{1}{J}}+{\displaystyle \frac{1}{3J^{}}}+{\displaystyle \frac{\gamma }{J^{}}}\right),`$ $`{\displaystyle \frac{1}{J^{(2)}}}={\displaystyle \frac{1}{8}}\left({\displaystyle \frac{1}{J}}+{\displaystyle \frac{5}{3J^{}}}+{\displaystyle \frac{\gamma }{J^{}}}\right).`$ (9) Comparing Eq. (4) with Eq. (7), we have the topological angle $`\theta =4\pi {\displaystyle \frac{J^{(0)}}{J^{(1)}}}=\pi +{\displaystyle \frac{\pi \gamma }{{\displaystyle \frac{J^{}}{J}}+{\displaystyle \frac{1}{3}}}},`$ (10) the spin wave velocity $`v={\displaystyle \frac{4aJ^{}}{\sqrt{\left({\displaystyle \frac{J^{}}{J}}+{\displaystyle \frac{1}{3}}\right)\left({\displaystyle \frac{J^{}}{J}}+3\right)\gamma ^2}}}`$ (11) and the coupling constant $`g={\displaystyle \frac{4\left({\displaystyle \frac{J^{}}{J}}+{\displaystyle \frac{1}{3}}\right)}{\sqrt{\left({\displaystyle \frac{J^{}}{J}}+{\displaystyle \frac{1}{3}}\right)\left({\displaystyle \frac{J^{}}{J}}+3\right)\gamma ^2}}}.`$ (12) These quantities are shown in Figs. 2, 3 and 4, as functions of $`J^{}/J`$ for several values of $`\gamma `$. ## 3 Phase diagram It is well known that an NLSM has gapless excitations if the topological angle $`\theta `$ is an odd-integer multiple of $`\pi `$. In the general periodic spin model, this condition is turned into the gapless equation $$\frac{2J^{(0)}}{J^{(1)}}=\frac{2l1}{2}$$ (13) with arbitrary integer $`l`$. In the present case of Eq. (10), it becomes of a simple form $`\gamma =2l({\displaystyle \frac{J^{}}{J}}+{\displaystyle \frac{1}{3}}).(l=1,0,1)`$ (14) The equation for each $`l`$ determines a phase boundary between gapful disordered phases. The restricted values of integer $`l`$ is due to the condition $`1<\gamma <1`$. In the symmetric case of $`\gamma =0`$, the system is gapless irrespective of the ratio $`J^{}/J`$, because it satisfies Eq. (14) with $`l=0`$. The phase diagram of the ground state is obtained by identifying the gapless lines (14) as the phase boundaries. We show the phase diagram in Fig. 5. There exist four phases A<sub>+</sub>, A<sub>-</sub>, B<sub>+</sub> and B<sub>-</sub> in the ($`J^{}/J`$, $`\gamma `$) parameter space. The phase structure is symmetric for $`\gamma \gamma `$ according to the symmetry of the Hamiltonian (1) with Eqs. (2) and (3). In phase B<sub>+</sub>, $`\gamma `$ is close to 1 and $`J_2`$ is selectively large. The ground state of B<sub>+</sub> may be explained by a VBS picture . That is, if we decompose each spin into spins with magnitude 1/2, the ground state is approximately a direct product of dimers. Two dimers in a unit cell are formed between 1/2-spins at the both sides of a coupling with exchange parameter $`J_2`$. Phase A<sub>+</sub> is of an intermediate character and spreads over almost all area in the parameter space. The ground state of A<sub>+</sub> does not seem to be explained by the conventional VBS picture. ## 4 Summary We have examined the 1/2-1-3/2-1 spin chain with a distortion of the exchange parameters by using the nonlinear $`\sigma `$ model method. We obtained the phase diagram of the ground state in the parameter space. There exist four gapful phases. In the case of no distortion, the system is always gapless and is on a phase boundary. When the distortion is strong, there are two phases to be explained by a VBS picture. It is a future work to explain the other phases by the singlet-cluster-solid (SCS) picture developed in Ref. .
warning/0003/cond-mat0003089.html
ar5iv
text
# Quantum Dot as Spin Filter and Spin Memory \[ ## Abstract We consider a quantum dot in the Coulomb blockade regime weakly coupled to current leads and show that in the presence of a magnetic field the dot acts as an efficient spin-filter (at the single-spin level) which produces a spin-polarized current. Conversely, if the leads are fully spin-polarized the up or down state of the spin on the dot results in a large sequential or small cotunneling current, and thus, together with ESR techniques, the setup can be operated as a single-spin memory. \] An increasing number of spin-related experiments show that the spin of the electron offers unique possibilities for finding novel mechanisms for information processing—most notably in quantum-confined semiconductors with unusually long spin dephasing times approaching microseconds , and where spins can be transported coherently over distances of up to 100 micrometers . Besides the intrinsic interest in spin-related phenomena, spin-based devices hold promises for future applications in conventional as well as in quantum computer hardware . One of the challenging problems for such applications is to obtain sufficient control over the spin dynamics in nanostructures. In the following we address this issue and propose a quantum-dot setup which can be either operated as a spin-filter (spin diode) to produce spin-polarized currents or as a device to detect and manipulate single-spin states (single-spin memory). Both effects occur at the single-spin level and thus represent the ultimate quantum limit of a spin-filter and spin-memory. In both cases, we will work in the Coulomb blockade regime and consider sequential and cotunneling processes. A new feature of our proposal is that the spin-degeneracy is lifted with different Zeeman splittings in the dot and in the leads which then results in Coulomb blockade peaks which are uniquely associated with a definite spin state on the dot. Formalism.– Our system consists of a quantum dot (QD) connected to two Fermi-liquid leads which are in equilibrium with reservoirs kept at the chemical potentials $`\mu _l`$, $`l=1,2`$, where outgoing currents can be measured, see Fig. 1. Using a standard tunneling Hamiltonian approach, we write for the full Hamiltonian $`H_0+H_T`$, where $`H_0=H_L+H_D`$ describes the leads and the dot, with $`H_D`$ including the charging and interaction energies of the dot electrons as well as the Zeeman energy $`g\mu _BB`$ of their spins in the presence of an external magnetic field $`𝐁=(0,0,B)`$, where $`g`$ is the effective g-factor. We concentrate first on unpolarized lead currents and assume that the Zeeman splitting in the leads is negligibly small compared to the one in the QD. This can be achieved e.g. by using InAs for the dot ($`g=15`$) attached to GaAs 2DEG leads ($`g=0.44`$), or by implanting a magnetic impurity (say Mn) inside a GaAs dot (again attached to GaAs 2DEG leads) with a strongly enhanced electron g-factor due to exchange splitting with the magnetic impurity . \[Below we will also consider the opposite situation with a fully spin polarized lead current, and a much smaller Zeeman splitting on the dot.\] The tunneling between leads and the QD is described by the perturbation $`H_T=_{l,k,p,s}t_{lp}c_{lks}^{}d_{ps}+\mathrm{h}.\mathrm{c}.`$, where $`d_{ps}`$ and $`c_{lks}`$ annihilate electrons with spin $`s`$ in the dot and in the $`l`$th lead, resp. While the orbital $`k`$-dependence of the tunneling amplitude $`t_{lp}`$ can be safely neglected, this is not the case in general for the QD orbital states $`p`$. From now on we concentrate on the Coulomb blockade (CB) regime, where the charge in the QD, $`\widehat{N}=_{p,s}d_{ps}^{}d_{ps}`$, is quantized, $`\mathrm{Tr}\rho \widehat{N}=N`$. Next, turning to the dynamics induced by $`H_T`$, we introduce the reduced density matrix for the dot, $`\rho _D=\mathrm{Tr}_L\rho `$, where $`\rho `$ is the full stationary density matrix, and $`\mathrm{Tr}_L`$ is the trace over the leads. To describe the stationary limit, we use a standard master equation approach formulated in terms of the dot eigenstates and eigenenergies, $`H_D|n=E_n|n`$, where $`n=(𝐧,N)`$. Denoting with $`\rho (n)=n|\rho _D|n`$ the stationary probability for the dot to be in the state $`|n`$, and with $`W(n^{},n)`$ the transition rates between $`n`$ and $`n^{}`$, the stationary master equation to be solved reads $`_n\left[W(n^{},n)\rho (n)W(n,n^{})\rho (n^{})\right]=0`$. The rates $`W`$ can be calculated in a standard “golden rule” approach where we go up to 2nd order in $`H_T`$, i.e. $`W=_lW_l+_{l^{},l}W_{l^{}l}`$, where $`W_lt^2`$ is the rate for a tunneling process of an electron from the $`l`$th lead to the dot and back, while $`W_{l^{}l}t^4`$ describes the simultaneous tunneling of two electrons from the lead $`l`$ to the dot and from the dot to the lead $`l^{}`$. Thus, two regimes of transport through the QD can be distinguished: Sequential tunneling (ST) and cotunneling (CT). The ST regime is at the degeneracy point, where $`\widehat{N}`$ fluctuates between $`N`$ and $`N^{}=N\pm 1`$, and 1st order transitions are allowed by energy conservation with the explicit rates $`W_l(n^{},n)`$ $`=`$ $`2\pi \nu [f_l(\mathrm{\Delta }_{n^{}n})|A_{lnn^{}}|^2\delta _{N^{},N+1}`$ (1) $`+`$ $`[1f_l(\mathrm{\Delta }_{nn^{}})]|A_{ln^{}n}|^2\delta _{N^{},N1}],`$ (2) where $`\nu =_k\delta (\epsilon _F\epsilon _k)`$ is the lead density-of-states per spin at the Fermi energy $`\epsilon _F`$, $`f_l(\epsilon )=[1+\mathrm{exp}((\epsilon \mu _l)/k_BT)]^1`$ the Fermi function at temperature $`T`$, $`\mathrm{\Delta }_{n^{}n}=E_n^{}E_n`$ is the level distance, and we have introduced the matrix elements $`A_{ln^{}n}=_{ps}t_{lps}n^{}|d_{ps}|n`$. In the ST regime the current through the QD can be written as $`I_s=\pm e_{n,n^{}}W_2(n^{},n)\rho (n)`$, where $`\pm `$ stands for $`N^{}=N1`$. We emphasize that the rates $`W(n,n^{})`$ and thus the current depend on the spin state of the dot electrons via $`n,n^{}`$. The ST current takes a particularly simple form if bias $`\mathrm{\Delta }\mu =\mu _1\mu _2>0`$ and temperature are small compared to the level distance on the dot (the case of interest here), $`\mathrm{\Delta }\mu ,k_BT<|\mathrm{\Delta }_{mn}|,m,n`$, and, thus only the lowest energy levels participate in the transport . The solution of the master equation gives in this case for the ST current $$I_s=\frac{e\gamma _1\gamma _2}{\gamma _1+\gamma _2}\left[f_1(\mathrm{\Delta }_{n^{}n})f_2(\mathrm{\Delta }_{n^{}n})\right],$$ (3) where $`\gamma _l=2\pi \nu |A_{lnn^{}}|^2`$ is the tunneling rate through the $`l`$th barrier. In the CT regime the only allowed processes are 2nd order transitions with initial and final electron number on the QD being equal, i.e. $`N=N^{}`$, and with the rate $`W_{l^{}l}(n^{},n)=2\pi \nu ^2{\displaystyle 𝑑\epsilon f_l(\epsilon )[1f_l^{}(\epsilon \mathrm{\Delta }_{n^{}n})]}`$ (4) $`\times \left|{\displaystyle \underset{n_1}{}}{\displaystyle \frac{A_{l^{}n^{}n_1}A_{lnn_1}^{}}{\mathrm{\Delta }_{nn_1}+\epsilon }}+{\displaystyle \underset{n_2}{}}{\displaystyle \frac{A_{l^{}n_2n}A_{ln_2n^{}}^{}}{\mathrm{\Delta }_{n^{}n_2}\epsilon }}\right|^2,`$ (5) where, $`N_1=N+1`$, and $`N_2=N1`$, and thus the two terms in Eq. (5) differ by the sequence of tunneling. Our regime of interest here is elastic CT where $`E_n^{}=E_n`$, which holds for $`|\mathrm{\Delta }_{mn}|>\mathrm{\Delta }\mu ,k_BT,mn`$. That means the system is always in the ground state with $`\rho (n)=1`$, and thus the CT current is given by $`I_c=eW_{21}(n,n)eW_{12}(n,n)`$. In particular, close to a ST resonance (but still in the CT regime) Eq. (5) considerably simplifies—only one term contributes—and for $`\mathrm{\Delta }\mu ,k_BT<|\mu \pm \mathrm{\Delta }_{nn_i}|`$, we obtain $$I_c=\frac{e}{2\pi }\frac{\gamma _1\gamma _2\mathrm{\Delta }\mu }{(\mu \pm \mathrm{\Delta }_{nn_i})^2},$$ (6) where $`+`$ stands for $`i=1`$, and $``$ for $`i=2`$, and $`\mu =(\mu _1+\mu _2)/2`$. From Eqs. (3) and (6) it follows that $`I_s\gamma _i`$, while $`I_c\gamma _i^2`$, and therefore $`I_cI_s`$. Thus, in the CB regime the current as a function of $`\mu `$ (or gate voltage) consists of resonant ST peaks, where $`\widehat{N}`$ on the QD fluctuates between $`N`$ and $`N\pm 1`$. The peaks are separated by plateaus, where $`N`$ is fixed, and where the (residual) current is due to CT. We note that the tunneling rates $`\gamma _l`$ depend on the tunneling path through the matrix elements $`A_{lmn}`$. In general, this can lead to a spin-dependence of the current, which, however, is difficult to measure. In contrast to this, we will show now that a much stronger spin dependence can come from the resonance character of the currents $`I_s`$ and $`I_c`$, when the position of a resonance (as function of gate voltage) depends on the spin orientation of the tunneling electron. To proceed we first specify the energy spectrum of the QD more precisely. In general, the determination of the spectrum of a QD is a complicated many-electron problem . However, it is known from experiment that especially away from orbital degeneracy points (which can be easily achieved by applying magnetic fields) the spectrum is formed mainly by single-particle levels, possibly slightly renormalized by exchange interaction . For a QD with $`N`$ odd there is one unpaired electron in one of the two lowest energy states, $`|`$ and $`|`$, with energies $`E_{}`$ and $`E_{}`$, which become Zeeman split due to a magnetic field $`B`$, $`\mathrm{\Delta }_z=|E_{}E_{}|=\mu _B|gB|`$. Let us assume that $`|`$ is the ground state, and set $`E_{}=0`$ for convenience. For $`N`$ even, the two topmost electrons (with the same orbital wave function), form a spin singlet, $`(||)/\sqrt{2}`$, with energy $`E_S`$. This is the ground state, while the other states, such as three triplets $`|T_+=|`$, $`|T_{}=|`$, and $`|T_0=(|+|)/\sqrt{2}`$ with energies $`E_{T_\pm }`$ and $`E_{T_0}`$ are excited states, because of higher (mostly) single-particle orbital energy. Sequential tunneling. First we consider the ST peak, which separates two plateaus with $`N`$ and $`N+1`$ electrons on the dot, where $`N`$ is odd (odd-to-even transition). In the regime $`E_{T_+}E_S,\mathrm{\Delta }_z>\mathrm{\Delta }\mu `$, $`k_BT`$, only ground-state transitions are allowed by energy conservation. The tunneling of spin-up electrons is blocked by energy conservation, i.e. $`I_s()=0`$, because it involves excited states $`|T_+`$ and $`|`$. The only possible process is tunneling of spin-down electrons as shown in Fig. 2, which leads to a spin-polarized current, $`I_s()`$, given by Eq. (3), where $`\mathrm{\Delta }_{n^{}n}=E_S>0`$ (since $`E_{}=0`$). Thus, we have $`I_s()/I_0=\theta (\mu _1E_S)\theta (\mu _2E_S),k_BT<\mathrm{\Delta }\mu ,`$ (7) $`I_s()/I_0={\displaystyle \frac{\mathrm{\Delta }\mu }{4k_BT}}\mathrm{cosh}^2\left[{\displaystyle \frac{E_S\mu }{2k_BT}}\right],k_BT>\mathrm{\Delta }\mu ,`$ (8) where $`I_0=e\gamma _1\gamma _2/(\gamma _1+\gamma _2)`$. Hence, in the specified regime the dot acts as spin filter through which only spin-down electrons can pass . Next, we consider the ST peak at the transition from even to odd, i.e. when $`N`$ is even. Then the current is given by Eq. (3) with $`\mathrm{\Delta }_{n^{}n}=E_S>0`$. The spin-down current is now blocked, $`I_s()=0`$, while spin-up electrons can pass through the QD, with the current $`I_s()`$ given by (7) and (8), where $`E_S`$ has to be replaced by $`E_S`$. Because this case is very similar to the previous one with $``$ replaced by $``$, we shall concentrate on the odd-to-even transition only. Next, we will demonstrate that although CT processes can in general lead to a leakage of unwanted current, this turns out to be a minor effect, and spin filtering works also in the CT regime. Cotunneling. Above or below a ST resonance, i.e. when $`E_S>\mu _{1,2}`$ or $`E_S<\mu _{1,2}`$, the system is in the CT regime. Close to this peak the main contribution to the transport is due to two processes (a) and (c) see Fig. 3, where the energy deficit of the virtual states, $`|\mu E_S|`$, is minimal. According to Eq. (6) we have $$I_c()=\frac{e}{2\pi }\frac{\gamma _1\gamma _2\mathrm{\Delta }\mu }{(\mu E_S)^2}.$$ (9) Thus, we expect the spin filtering of down electrons to work even in the CT regime close to the resonance. However, there are additional CT processes, (b) and (d), which involve tunneling of spin-up electrons and lead to a leakage of up-spin. If $`N`$ is odd (below the resonance), the dot is initially in its ground state $`()`$, and an incoming spin-up electron can only form a virtual triplet state $`|T_+`$ (process (b) in Fig. 3). This process contributes to the rate (3) with an energy deficit $`E_{T_+}\mu `$, so that for the efficiency of spin filtering \[defined as $`I()/I()`$\] we obtain in this regime, $$I_c()/I_c()\left(1+\frac{E_{T_+}E_S}{E_S\mu }\right)^2,N\text{ odd.}$$ (10) Above the resonance, i.e. when $`N`$ is even and the ground state is the spin-singlet $`|S`$, the tunneling of spin-up electrons occurs via the virtual spin-down state (process (d) in Fig. 3) with an energy deficit $`\mathrm{\Delta }_z+\mu E_S`$, which has to be compared to the energy deficit $`\mu E_S`$ of the main process (c). Thus, we obtain for the efficiency of the spin filtering in the CT regime $$I_c()/I_c()\left(1+\frac{\mathrm{\Delta }_z}{\mu E_S}\right)^2,N\text{ even.}$$ (11) We see that in both cases, above and below the resonance, the efficiency can be made large by tuning the gate voltage to the resonance, i.e. $`|\mu E_S|0`$. Eventually, the system goes to the ST regime, $`|\mu E_S|k_BT,\mathrm{\Delta }\mu `$. Now, using Eqs. (6), (7), and (8) we can estimate the efficiency of spin filtering in the ST regime, $$I_s()/I_c()\frac{\mathrm{\Delta }_z^2}{(\gamma _1+\gamma _2)\mathrm{max}\{k_BT,\mathrm{\Delta }\mu \}},$$ (12) where we have assumed that $`\mathrm{\Delta }_z<|E_{T_+}E_S|`$. In the ST regime considered here we have $`\gamma _i<k_BT,\mathrm{\Delta }\mu `$ . Therefore, if the requirement $`k_BT,\mathrm{\Delta }\mu <\mathrm{\Delta }_z`$ is satisfied, filtering of spin-down electrons in the ST regime is very effecient. In the quantum regime, $`\gamma _i>k_BT,\mathrm{\Delta }\mu `$, tunneling occurs as Breit-Wigner resonance, and $`\mathrm{max}\{k_BT,\mathrm{\Delta }\mu \}`$ in Eq. (12) has to be replaced by $`\gamma _i`$. Finally, we note that the spin polarization of the transmitted current oscillates between up and down as we change the number of dot electrons $`N`$ one by one. The functionality of the spin filter can be tested e.g. with the use of a p-i-n diode which is placed in the outgoing lead 2. Via excitonic photoluminescence, the diode transforms the spin polarized electrons (entering lead 2) into correspondingly circularly polarized photons which can then be detected. Spin memory. We consider now the opposite case where the current in the leads is fully spin polarized. Recent experiments have demonstrated that fully spin-polarized carriers can be tunnel-injected from a spin-polarized magnetic semiconductor (III-V or II-VI materials) with large effective g-factor into an unpolarized GaAs system . Another possibility would be to work in the quantum Hall regime where spin-polarized edge states are coupled to a quantum dot. To be specific, we consider the case where $`E_{T_+}E_s+\mathrm{\Delta }_z>\mathrm{\Delta }\mu ,k_BT`$ with $`E_{T_+}>E_s`$ ($`\mathrm{\Delta }_z>k_BT`$ is not necessary as long as the spin relaxation time is longer than the measurement time, see below). We assume that the spin polarization of both leads is, say, up and $`N`$ is odd. There are now two cases for the current, $`I^{}`$ or $`I^{}`$, corresponding to a spin up or down on the QD. First, we assume the QD to be in the ground state with its topmost electron-spin pointing up. According to previous analysis \[see paragraph before Eqs. (7,8)\], the ST current vanishes, i.e. $`I_s^{}=0`$, since the tunneling into the level $`E_{T_+}`$ (and higher levels) is blocked by energy conservation, while the tunneling into $`E_s`$ is blocked by spin conservation (the leads can provide and take up only electrons with spin up). However, there is again a small CT current, $`I_c^{}`$, which is given by Eq. (9). Now we compare this to the second case where the topmost dot-spin is down with additional Zeeman energy $`\mathrm{\Delta }_z>0`$. Here, the ST current $`I_s^{}`$ is finite, and given by Eqs. (7, 8) with $`E_s`$ replaced by $`E_s\mathrm{\Delta }_z`$. In the limit $`E_s>\mathrm{\Delta }_z`$, we get $`I_s^{}=I_s()`$, and thus the ratio $`I_s^{}/I_c^{}`$ is again given by Eq. (12). Hence, we see that the dot with its spin up transmits a much smaller current than the dot with spin down. This fact allows the read-out of the spin-state of the (topmost) dot-electron by comparing the measured currents. Furthermore, the spin state of the QD can be changed (“read-in”) by ESR techniques, i.e. by applying a pulse of an ac magnetic field (perpendicular to B) with resonance frequency $`\omega =\mathrm{\Delta }_z`$. Thus the proposed setup functions as a single-spin memory with read-in and read-out capabilities, the relaxation time of the memory given by the spin relaxation time $`\tau _S`$ on the QD (which can be expected to exceed 100’s of nanoseconds). We note that this $`\tau _S`$ can be easily measured since it is the time during which $`I_s^{}`$ is finite before it strongly reduces to $`I_c^{}`$. Finally, this scheme can be upscaled: In an array of such QDs where each dot separately is attached to in- and outgoing leads (for read-out) we can switch the spin-state of each dot individually by locally controlling the Zeeman splitting $`\mathrm{\Delta }_z`$. This can be done e.g. by applying a gate voltage on a particular dot that pushes the wave function of the dot-electrons into a region of, say, higher effective g-factor (the induced level shift in the QD can be compensated for by the chemical potentials). In conclusion, we have shown that quantum dots in the Coulomb blockade regime and attached to leads can be operated as efficient spin filters at the single-spin level. Conversely, if the leads are spin-polarized the spin state of the quantum dot can be read-out by a traversing current which is (nearly) blocked for one spin state while it is unblocked for the opposite spin state. Acknowledgements. This work has been supported by the Swiss NSF.
warning/0003/cond-mat0003209.html
ar5iv
text
# Localization-induced Griffiths phase of disordered Anderson lattices \[ ## Abstract We demonstrate that local density of states fluctuations in disordered Anderson lattice models universally lead to the emergence of non-Fermi liquid (NFL) behavior. The NFL regime appears at moderate disorder ($`W=W_c`$) and is characterized by power-law anomalies, e. g. $`C/T1/T^{(1\alpha )}`$, where the exponent $`\alpha `$ varies continuously with disorder, as in other Griffiths phases. This Griffiths phase is not associated with the proximity to any magnetic ordering, but reflects the approach to a disorder-driven metal-insulator transition (MIT). Remarkably, the MIT takes place only at much larger disorder $`W_{MIT}12W_c`$, resulting in an extraordinarily robust NFL metallic phase. \] The nature of the non-Fermi liquid (NFL) behavior observed in several heavy fermion compounds remains largely unresolved. In the cleaner systems, the proximity to a quantum critical point seems to be at the origin of many of the observed properties . Exotic impurity models cannot be discarded, though their behavior in concentrated systems remains ill understood. In other compounds, non-stoichiometry has prompted the investigation of disorder-based mechanisms. A phenomenological “Kondo disorder” model (KDM), describing a broad distribution of Kondo temperatures $`T_K`$, has been successfully applied to several of these systems. Alternatively, the formation of large clusters of magnetically ordered material within the disordered phase has also been proposed. Both scenarios lead to a wide distribution of energy scales, giving rise to similar thermodynamic anomalies and NMR response. In addition, the predictions of the KDM prove to be consistent with a number of other experiments, including optical conductivity, magnetoresistance and dynamic neutron scattering measurements. Despite these successes, a number of basic questions remain unresolved, including: (1) What is the microscopic origin of the ubiquitous power law (or logarithmic) anomalies? (2) Can a model calculation be done, which can produce these power laws in a universal fashion? (3) Are these properties tied to the proximity to a quantum phase transition, and if so, which one? (4) How robust is the anomalous behavior with respect to the variation of materials parameters? Within our model, all these questions find clear-cut and physically transparent answers: (i) The anomalies can be ascribed to a power law distribution of $`T_K`$’s, whose exponent varies continuously with disorder strength. The resulting NFL behavior, e. g. $`\gamma =C/T1/T^{(1\alpha )}`$, $`\alpha <1`$ sets in for relatively weak randomness, irrespective of the detailed model for disorder. This should be contrasted with the KDM , where the occurence of NFL behavior requires fine-tuning. (ii) We find universal behavior reflecting the nonlocal, many-body processes associated with Anderson localization effects in the presence of strong electron correlations. (iii) for stronger disorder, the NFL metallic behavior persists over a surprisingly large interval before a disorder-driven MIT is reached. This novel Griffiths phase is a manifestation of quantum critical behavior associated with the approach to a disorder-driven metal-insulator transition and does not require the proximity of any magnetically ordered phase. We consider a disordered infinite-U Anderson lattice Hamiltonian $`H`$ $`=`$ $`{\displaystyle \underset{ij\sigma }{}}\left(t_{ij}+\epsilon _i\delta _{ij}\right)c_{i\sigma }^{}c_{j\sigma }^{}+{\displaystyle \underset{j\sigma }{}}E_{fj}f_{j\sigma }^{}f_{j\sigma }^{}`$ (1) $`+`$ $`{\displaystyle \underset{j\sigma }{}}V_j(c_{j\sigma }^{}f_{j\sigma }^{}+\mathrm{H}.\mathrm{c}.),`$ (2) in usual notation. The infinite-U constraint at each f-orbital is assumed ($`n_j^f=_\sigma f_{j\sigma }^{}f_{j\sigma }^{}1`$). We have studied different types of disorder, including randomness in the conduction electron site energies $`\epsilon _i`$, the f-electron energies $`E_{fj}`$, as well as the hybridization $`V_j`$. Within our approach, we find that most of our conclusions remain valid for any specific form of disorder, indicating robust and universal behavior. We treat the above Hamiltonian within the recently proposed statistical dynamical mean field theory. This approach reduces to the usual dynamical mean field theory in the limit $`z\mathrm{}`$ (with $`t_{jk}t/\sqrt{z}`$), but unlike the latter, it incorporates Anderson localization effects. As a result, the spectral function of the local bath “seen” by each impurity has strong spatial fluctuations and contains information coming from sites which are many lattice parameters away. Physically, the fluctuations of the conduction electron wave-functions lead to the distribution of Kondo temperatures, which in turn creates a renormalized effective disorder seen by the conduction electrons. This nonlocal feedback mechanism results in the universal form of all the relevant distribution functions that we find. The simplest model for incorporating localization effects is obtained by focusing on a Bethe lattice of coordination $`z`$ (with nearest neighbor hopping $`t`$, used as unit of energy). The resulting set of self-consistent stochastic equations is governed by the local actions $`S_{\mathrm{eff}}^{(j)}`$ $`=`$ $`{\displaystyle _0^\beta }𝑑\tau {\displaystyle \underset{\sigma }{}}f_{j,\sigma }^{}(\tau )\left(_\tau +E_{fj}\right)f_{j,\sigma }(\tau )+`$ (5) $`{\displaystyle _0^\beta }𝑑\tau {\displaystyle _0^\beta }𝑑\tau ^{}{\displaystyle \underset{\sigma }{}}f_{j,\sigma }^{}(\tau )\mathrm{\Delta }_j(\tau \tau ^{})f_{j,\sigma }(\tau ^{});`$ $`\mathrm{\Delta }_j(\omega )={\displaystyle \frac{V_j^2}{\omega \epsilon _j_{k=1}^{z1}t_{jk}^2G_{ck}^{(j)}(\omega )}}.`$ Here, $`G_{ck}^{(j)}(\omega )`$ is the local c-electron Green’s function on site $`k`$ with the nearest neighbor site $`j`$ removed. It is determined recursively from $`G_{cj}^{(i)(1)}(\omega )`$ $`=`$ $`\omega \epsilon _j{\displaystyle \underset{k=1}{\overset{z1}{}}}t_{jk}^2G_{ck}^{(j)}(\omega )\mathrm{\Phi }_j(\omega ),`$ (6) $`\mathrm{\Phi }_j(\omega )`$ $`=`$ $`{\displaystyle \frac{V_j^2}{\omega E_{fj}\mathrm{\Sigma }_{fj}(\omega )}}.`$ (7) The local self-energy $`\mathrm{\Sigma }_{fj}(\omega )`$ is obtained from the solution of the effective action (5). In order to solve the impurity problems, we have used the large-N mean-field theory at $`T=0`$. We have solved Eqs. (5-7) numerically by sampling. In implementing this procedure, we have carried out large-scale simulations for $`z=3`$, with ensembles containing up to $`N_s=200`$ sites, and frequency meshes containing up to $`N_{omega}=8,000`$ frequencies. The numerical integrations needed to solve the impurity problems have been done by a combination of spline interpolations and adaptive quadrature routines. These careful numerics have made it possible to obtain Kondo temperatures spanning fifteen orders of magnitude, which was crucial in order to examine the long tails of the relevant distribution functions. One of the greatest advantages of our approach is its ability to focus on full distribution functions, which is essential for characterizing any Griffiths phase. Some typical results are presented in Fig. 1, where we show the evolution of the distribution of local Kondo temperatures as a function of disorder, from which one computes the overall response of the lattice system (See the discussion in the first ref. of ). We find that (Fig. 1(a)) the distribution has a universal log-normal form for weak disorder. We have verified that such a log-normal behavior is obtained irrespective of the type and shape of the bare disorder distribution, as long as it is not too large. As the disorder is increased, the distribution $`P(T_K)`$ no longer retains its log-normal form. Instead, a long tail emerges on the low-$`T_K`$ side, with a power law asymptotic form (Fig. 1(b)) $$P(T_K)T_K^{(\alpha 1)}.$$ (8) The exponent $`\alpha `$ varies continuously with disorder, as seen on a plot of $`\mathrm{log}(P(\mathrm{log}(T_K)))`$ in Fig. 2. Note that the value $`\alpha =1`$, (Figs. 1(b) and 2) with $`P(T_K)\mathrm{const}.`$, corresponds to the condition for Marginal Fermi Liquid behavior observed in some Kondo alloys, with logarithmically divergent magnetic susceptibility $`\chi (T)`$ and specific heat coefficient $`\gamma `$. This divergent behavior becomes more singular as the disorder is increased past this marginal case. For example, if we use the simple Wilson interpolation formula for $`\chi (T)`$ $$\chi (T)_0^\mathrm{\Lambda }\frac{T_K^{(\alpha 1)}}{T+aT_K}𝑑T_K\frac{1}{T^{(1\alpha )}}.$$ (9) We thus have power law divergences with exponents which vary continuously with the disorder strength. If we take $`t10^4K`$, this should be observed below a few tens of Kelvin. Such generic behavior has been fitted to some NFL compounds. Besides, $`\chi (0)`$ will diverge at a critical disorder strength ($`W_c1`$ in Fig. 2) $`\chi (0)`$ $``$ $`{\displaystyle \frac{P(T_K)}{T_K}𝑑T_K}{\displaystyle _0^\mathrm{\Lambda }}T_K^{(\alpha 2)}𝑑T_K`$ (10) $``$ $`{\displaystyle \frac{1}{\alpha 1}}{\displaystyle \frac{1}{W_cW}},`$ (11) with a similar result for $`\gamma `$. Note, however, that other higher order correlation functions, such as the non-linear susceptibility $`\chi _3(0)`$, which probes higher negative moments of the distribution ($`\chi _3(0)1/T_K^3`$), will begin to diverge at different critical values of disorder, $$\chi _3(0)\frac{1}{\alpha 3}\frac{1}{W_{c3}W},$$ (12) where $`W_{c3}0.66`$ for the parameters of Fig. 2. This general behavior is characteristic of Griffiths phases and should not be confused with a true phase transition. The system should be viewed as a disordered metal with embedded clusters of Anderson insulators. It is precisely these poor conducting regions, with depleted densities of states, which give rise to imperfectly quenched spins and the corresponding singular thermodynamic properties. We should also stress that the main mechanism that dominates the Griffiths phase is qualitatively different from the one in the KDM. There, $`T_K`$ fluctuations were simply caused by the distribution of local parameters ($`V_j`$, $`E_{fj}`$) and the conduction electron DOS does not fluctuate. By contrast, in the present treatment, fluctuations in the latter are dominant. To illustrate this, all the results we present are obtained for a model with conduction band disorder only, although similar results follow for any form of disorder. We stress that, in a KDM treatment of this case, $`T_K`$ fluctuations are severely limited. Here, however, $`T_K`$ fluctuations are enhanced by the fluctuations in the local conduction DOS, reflecting the localization effects and the approach to a disorder-driven MIT. To confirm this picture, we examine the localization properties of the conduction electrons. We focus on the typical DOS $`\rho _{typ}=\mathrm{exp}\{<\mathrm{ln}\rho _j>\}`$; $`\rho _j=(1/\pi )ImG_{cj}(\omega =0)`$, as shown in Fig. 3. This quantity vanishes at any disorder-driven MIT, and thus can serve as an order parameter for localization. Remarkably, we find a strong decrease of this quantity upon entering the Griffiths phase ($`W/t1`$), reflecting the strongly enhanced conduction electron scattering due to Kondo disorder. Yet, the actual localization transition, where the typical DOS vanishes, occurs only at much larger disorder ($`W/t12`$). This results in a very extended NFL metallic region, where the thermodynamics is singular, and the conduction electrons are almost, but not completely localized. This dramatic effect has a simple physical origin. Consider the distribution of the effective scattering potentials of the conduction electrons $`\mathrm{\Phi }_j(\omega =0)`$ (see Eq. (5)) introduced by the f-sites. Note that $`\mathrm{\Phi }_j(\omega =0)=Z_jV^2/\stackrel{~}{\epsilon }_{fj}`$, where $`Z_j`$ is the quasiparticle weight and $`\stackrel{~}{\epsilon }_{fj}`$ the (renormalized) energy of the Kondo resonance at site $`j`$. For sufficient disorder, the Kondo resonances are randomly shifted up or down in energy, giving rise to $`\mathrm{\Phi }_j`$’s that can be random in magnitude but also in sign. The resulting distribution for the inverse quantity $`\mathrm{\Phi }_j^1`$ is shown in Fig. 4 and is found to broaden with disorder. For $`W/t1.5`$, a finite density of $`\mathrm{\Phi }_j^1=0+`$ (i. e. $`\mathrm{\Phi }_j=+\mathrm{}`$) sites emerges. This is crucial, since the corresponding f-sites act as unitary scatterers (US’s), characterized by a maximally allowed scattering phase shift ($`\delta =\pi /2`$) for the conduction electrons. If all the f-sites were US’s, the system would be a Kondo insulator. The presence of a finite fraction of US’s should be viewed as the emergence of droplets of a Kondo insulator within the heavy metal. Interestingly, at stronger disorder ($`W/t>4`$) the distribution of $`\mathrm{\Phi }_j^1(0)`$ continues to broaden, leading to a decrease in the number of US’s. This is illustrated by plotting $`P(\mathrm{\Phi }_j^1=0)`$ in the inset of Fig 4. In this regime, while the bare disorder increases, the effective disorder produced by the f-sites is reduced, stabilizing the almost localized metallic phase. This mechanism may be at the origin of the puzzling behavior of materials such as $`\mathrm{SmB}_6`$, where the low temperature resistivity remains anomalously large yet finite over a broad range of parameters. Finally, we note that a similar NFL phase was identified in a study of the Mott-Anderson transition in the disordered Hubbard model. We have re-examined this system, and concluded that the NFL behavior should be attributed to a related Griffiths phase rather than a separate thermodynamic phase of the system. Despite the similarities, several features prove dramatically different. For Hubbard models, the emergence of NFL behavior does not have a dramatic effect on the conduction electrons and no US’s emerge. This observation may explain the strong correlation between thermodynamic and transport anomalies in Kondo alloys, but not in doped semiconductors. In the latter materials, the thermodynamics is still singular close to the MIT, while transport remains more conventional. It would be of particular interest if it could be tested experimentally whether these localization effects are responsible for the observed NFL behavior of disordered heavy fermion systems. A scanning tunneling microscopy study might be able to detect the predicted insulator droplets. In order to distinguish this from the magnetic Griffiths phase scenario, a systematic study of systems with comparable amounts of disorder but different magnetic character would be useful. Besides, since the present theory relies very little on intersite magnetic correlations, a determination of the typical size of the relevant magnetic clusters could also serve as a test. In summary, we have investigated and solved a microscopic model for disordered Anderson lattices that displays an unprecedented sensitivity to disorder, leading to localization-induced non-Fermi liquid behavior. Our results demonstrate that a well defined Griffiths phase can exist, which is not restricted to the vicinity of any magnetic ordering and yet seems to be consistent with most puzzling features of disordered heavy fermion systems and Kondo alloys. We acknowledge useful discussions with M. C. Aronson, D. L. Cox, G. Kotliar, D. E. MacLaughlin, A. J. Millis, and G. Zarand. This work was supported by FAPESP and CNPq (EM), NSF grant DMR-9974311 and the Alfred P. Sloan Foundation (VD).
warning/0003/hep-ph0003039.html
ar5iv
text
# 1 Introduction ## 1 Introduction It is hard to imagine modern particles physics without such fundamental notions as, for instance, the phase transitions, topological defects, taken from statistical physics. This extremely fruitful connection among two branches of physics based on the euclidean postulate : the formulae of particle physics are coincide with corresponding formulae of statistical physics if the transformation $`tit`$ is applied. But this coincidence exist iff the media is equilibrium only, since the time order of physical process becomes lost after the transition to imaginary time $`it`$. So, the particles static properties only can be considered by euclidean field theories. The euclidean postulate does not ‘work’ for arbitrary element of $`S`$-matrix and, by this reason, there is not, at first glance, general connection between particles and statistical physics. Our aim is demonstrate this connection considering the multiple production example, staying in the real-time theory frame. The multiple production is a typical dissipative process of the incident kinetic energies transition into the energies (masses) of produced particles. This is the nonequilibrium process and the fluctuations, generally speaking, may be high in it. Experimental data confirms this general expectation at the mean multiplicities region, when $`n\overline{n}`$ . Considering multiple production we would like to note firstly that the mean multiplicity $`\overline{n}`$ of hadrons for modern accelerator energies ($``$10 Tev) is large $`\overline{n}(s)`$100. So, it is practically impossible to describe the system with $`N=3\overline{n}10`$ 300 degrees of freedom using ordinary methods. Secondly, it is natural to assume that the entropy $`𝒮`$ tends to maximum with rising multiplicity $`n`$ and reach the maximum at $`nn_{max}\sqrt{s}`$, since the dissipation take place in the vacuum (presumably with zero energy density)<sup>2</sup><sup>2</sup>2This consideration lie in the basis of earliest Fermi-Landau ‘statistical’ model of hadrons multiple production.. But the experiment shows that at high energies $`n\overline{n}(s)\mathrm{ln}^2s<<n_{max}(s)`$ are essential. This means that there is not total dissipation of incident energy in considered thermalization process . Absence of thermalization may be a consequence of hidden conservation laws . We would like to adopt following fundamental principle of nonequilibrium statistics introduced by N.N.Bogolyubov . It is natural to assume that the system evaluate to the equilibrium in such a way that the ‘nonequilibrium’ fluctuations in it should tend to zero. In the frame of Bogolyubov’s principle the quantitative measure of ‘nonequilibrium’ fluctuation is the mean value of correlation functions and, therefore, this quantities should tend to zero when the media tends to equilibrium. In our interpretation the Bogolyubov’s correlations relaxation principle means following. So, for nonequilibrium state presence ‘nonequilibrium’ fluctuations in the form of the macroscopic flow of, for instance, energy $`\epsilon `$ is natural. Then the mean value of $`m`$-point correlation functions $`K_m`$ can not be small as the consequence of macroscopic flow. But in vicinity of equilibrium the macroscopic flows should relax and, accordingly, the $`mean`$ value of correlation functions should be small, $`K_m0`$. To characterize the equilibrium one may consider also the particles, charge, spin, etc. densities macroscopic flows and theirs relaxation. We would like to show in result that the correlations relaxation principle leads to the quantitative connection with real time thermodynamics of Schwinger-Keldysh type<sup>3</sup><sup>3</sup>3Last one includes the nonequilibrium thermodynamics also. . Just for this purpose the generating functionals (GF) method of Bogolyubov will be used since it allows to find the quantitative connections, where the euclidean postulate does not applicable. We will use more natural for particles physics microcanonical formalism. In this formalism the thermodynamical ‘rough’ variables are introduced as the Lagrange multipliers of corresponding conservation laws. Theirs physical meaning are defined by corresponding equations of state. So, if the fluctuations in vicinity of solutions of corresponding equations are Gaussian then one can use this variables for description of the system. Corresponding condition is the Bogolyubov’s correlations relaxation condition. Formally, the generating functions method presents the integral transformation to new variables. One can choose them as the ‘rough’ thermodynamical variables. To describe the far from equilibrium system we will introduce the ‘local equilibrium hypothesis’. In its frame the preequilibrium state consist from equilibrium domains. In this case new variables should depend on the coordinates of domain and, in result, we are forced to use the generating functionals (GF) formalism. We will consider two example to illustrate effectiveness of the GF method. In Sec.2 we will consider the transformation (multiplicity $`n`$ activity $`z`$) to show the origin of the Koba-Nielsen-Olesen scaling (KNO-scaling)<sup>4</sup><sup>4</sup>4In privet discussion with one of the authors (A.S.) at summer of 1973 Z.Koba noted that the main reason of investigation leading to the KNO-scaling was just the GF method of N.N.Bogolyubov.. In Sec.3 we will investigate a possibility of temperature description of the multiple production processes. We will consider for this purpose the transformation (particles energies $`\epsilon `$ temperature $`1/\beta `$) to find the $`S`$-matrix interpretation of thermodynamics. It will be shown that this interpretation would be rightful iff the correlations are relax. In Sec.4 we will use this interpretation to formulate the perturbation theory in the case when $`\beta `$ and $`z`$ are local coordinates of temperature $`(x,t)`$ . One can use this closed form of perturbation theory for description of nonequilibrium media (in kinetic phase) and for description of the multiple production process as well. ## 2 KNO-scaling We would like start from note that the generating functions method allows connect inclusive spectra $`f_k`$ and exclusive cross sections $`\sigma _n(s)`$. One can use for this purpose the normalization condition: $$\overline{f}_k\sigma _{tot}𝑑\omega _k(q)f_k(q_1,q_2,\mathrm{},q_k)=\underset{n=k}{}\frac{n!}{(nk)!}\sigma _n,\overline{f}_k0\mathrm{}k>n_{max},$$ (2.1) where, as usual, $$d\omega _k(q)=\underset{i=1}{\overset{k}{}}d^3q_i/(2\pi )^32\epsilon (q_i),\epsilon (q)=\sqrt{q^2+m^2}$$ , is the Lorentz-covariant element of phase space. Eq.(2.1) can be considered as the set of coupled equations for $`\sigma _n`$. One may multiply both sides of (2.1) on $`(z1)^k/k!`$ and sum over $`k`$ to solve them. We will see that this is equivalent of introduction of ‘big partition function’ $`\mathrm{\Xi }(z)`$, where $`z`$ is the ‘activity’: the chemical potential $`\mu \mathrm{ln}z`$. We will find in result of summation over $`k`$ that $$\mathrm{\Xi }(z)\underset{k}{}\frac{(z1)^k}{k!}\overline{f}_k=\underset{n}{}z^n\frac{\sigma _n}{\sigma _{tot}}.$$ (2.2) Then, assuming that $`\mathrm{\Xi }(z)`$ is known, $$\sigma _n=\sigma _{tot}\frac{1}{2\pi i}_C\frac{dz}{z^{n+1}}\mathrm{\Xi }(z),$$ (2.3) where the closed contour $`C`$ includes point $`z=0`$. Here $`\mathrm{\Xi }(z)`$ is defined by left hand side of (2.2) and is the generating function of $`\sigma _n`$. The coefficients $`C_m`$ in decomposition: $$\mathrm{ln}\mathrm{\Xi }(z)=\underset{m}{}\frac{(z1)^m}{m!}C_m.$$ (2.4) are the (binomial) correlators. Indeed, $$C_1=\overline{f}_1=\overline{n},C_2=\overline{f}_2\left\{\overline{f}_1\right\}^2,C_3=\overline{f}_33\overline{f}_2\left\{\overline{f}_1\right\}^2+2\left\{\overline{f}_1\right\}^3$$ (2.5) an so on. If $`C_m=0`$, $`m>1`$, then $`\sigma _n`$ is described by Poisson formulae: $$\sigma _n=\sigma _{tot}e^{\overline{n}}\frac{(\overline{n})^n}{n!}.$$ (2.6) It corresponds to the case of absence of correlations. Let us consider more week assumption: $$C_m(s)=\gamma _m\left(C_1(s)\right)^m,$$ (2.7) where $`\gamma _m`$ is the energy independent constant. Then $$\mathrm{ln}\mathrm{\Xi }(z,s)=\underset{m=1}{}\frac{\gamma _m}{m!}\{(z1)\overline{n}(s)\}^m.$$ (2.8) To find consequences of this assumption let us find the mostly probable values of $`z`$. The equation: $$n=z\frac{}{z}\mathrm{ln}\mathrm{\Xi }(z,s)$$ (2.9) has increasing with $`n`$ solutions $`\overline{z}(n,s)`$ since $`\mathrm{\Xi }(z,s)`$ is the increasing function of $`z`$, iff $`\mathrm{\Xi }(z,s)`$ is the nonsingular at finite $`z`$ function. Last condition has deep physical meaning and practically assumes that absence of first order phase transition . Let us introduce new variable: $$\lambda =(z1)\overline{n}(s).$$ (2.10) Corresponding eq.(2.9) looks as follows: $$\frac{n}{\overline{n}(s)}=\left(1+\frac{\lambda }{\overline{n}(s)}\right)\frac{}{\lambda }\mathrm{ln}\mathrm{\Xi }(\lambda ).$$ (2.11) So, with $`O(\lambda /\overline{n}(s))`$ accuracy, one can assume that $$\lambda \lambda _c(n/\overline{n}(s)).$$ (2.12) are essential. It follows from this estimation that such scaling dependence is rightful at least in the neighborhood of $`z=1`$, i.e. in vicinity of main contributions into $`\sigma _{tot}`$. This gives: $$\overline{n}(s)\sigma _n(s)=\sigma _{tot}(s)\psi (n/\overline{n}(s)),$$ (2.13) where $$\psi (n/\overline{n}(s))\mathrm{\Xi }(\lambda _c(n/\overline{n}(s)))\mathrm{exp}\{n/\overline{n}(s)\lambda _c(n/\overline{n}(s))\}O(e^n)$$ (2.14) is the unknown function. The asymptotic estimation follows from the fact that $`\lambda _c=\lambda _c(n/\overline{n}(s))`$ should be, as follows from nonsingularity of $`\mathrm{\Xi }(z)`$, nondecreasing function of $`n`$. The estimation (2.12) is rightful at least at $`s\mathrm{}`$. The range validity of $`n`$, where solution of (2.12) is acceptable depends from exact form of $`\mathrm{\Xi }(z)`$. Indeed, if $`\mathrm{ln}\mathrm{\Xi }(z)\mathrm{exp}\{\gamma \lambda (z)\}`$, $`\gamma =const>0`$, then (2.12) is rightful at all values of $`n`$ and it is enough to have the condition $`s\mathrm{}`$. But if $`\mathrm{ln}\mathrm{\Xi }(z)(1+a\lambda (z))^\gamma `$, $`\gamma =const>0`$, then (2.12) is acceptable iff $`n<<\overline{n}^2(s)`$. Representation (2.13) shows that just $`\overline{n}(s)`$ is the natural scale of multiplicity $`n`$ . This representation was offered firstly as a reaction on the so called Feynman scaling for inclusive cross section: $$f_k(q_1,q_2,\mathrm{},q_k)\underset{i=1}{\overset{k}{}}\frac{1}{\epsilon (q_i)}.$$ (2.15) As follows from estimation (2.14), the limiting KNO prediction assumes that $`\sigma _n=O(e^n)`$. In this regime $`\mathrm{\Xi }(z,s)`$ should be singular at $`z=z_c(s)>1`$. The normalization condition $$\frac{\mathrm{\Xi }(z,s)}{z}|_{z=1}=\overline{n}(s)$$ gives: $`z_c(s)=1+\gamma /\overline{n}(s)`$, where $`\gamma >0`$ is the constant. Note, such behavior of big partition function $`\mathrm{\Xi }(z,s)`$ is natural for stationary Markovian processes described by logistic equations . In the field theory such equation describes the QCD jets . It is known that at $`Tevatron`$ energies the mean hadrons multiplicity rise with transverse momentum. The associated mean multiplicity is $$C_1(q_{tr})=\overline{n}(q_{tr})=\frac{_nnd\sigma _n/dq_{tr}}{_nd\sigma _n/dq_{tr}}.$$ So, if $$C_m(q_{tr})=\gamma _m\left(C_1(q_{tr})\right)^m:f_k(q_1,q_2,\mathrm{},q_k)\underset{i=1}{\overset{k}{}}\frac{1}{\epsilon (q_i)}\mathrm{\Omega }(q_{tr}),$$ then: $$\overline{n}(q_{tr})\frac{d\sigma _n/dq_{tr}}{_nd\sigma _n/dq_{tr}}=\mathrm{\Psi }(n/\overline{n}(q_{tr})).$$ This prediction is in good agreement with experiment . ## 3 Temperature description By definition, $$\sigma _n^{ab}(s)=𝑑\omega _n(q)\delta (q_a+q_b\underset{i=1}{\overset{n}{}}q_i)|A_n^{ab}|^2,$$ (3.1) where $`A_n^{ab}`$ is the amplitude of $`n`$ creation at interaction of particles $`a`$ and $`b`$. Considering Fourier transform of energy-momentum conservation $`\delta `$-function one can introduce the generating function $`\rho _n`$ . We may find in result that $`\sigma _n`$ is defined by equality: $$\sigma _n(s)=_i\mathrm{}^{+i\mathrm{}}\frac{d\beta }{2\pi }e^{\beta \sqrt{s}}\rho _n(\beta ),$$ (3.2) where $$\rho _n(\beta )=\left\{\underset{i=1}{\overset{n}{}}\frac{d^3q_ie^{\beta \epsilon (q_i)}}{(2\pi )^32\epsilon (q_i)}\right\}|A_n^{ab}|^2.$$ (3.3) The mostly probable value of $`\beta `$ in is defined by equation of state: $$\sqrt{s}=\frac{}{\beta }\mathrm{ln}\rho _n(\beta ).$$ (3.4) Let us consider the simplest example of noninteracting particles: $$\rho _n(\beta )=\left\{2\pi mK_1(\beta m)/\beta \right\}^n,$$ where $`K_1`$ is the Bessel function. Inserting this expression into (3.4) we can find that in the nonrelativistic case ($`nn_{max}`$) $$\beta _c=\frac{3}{2}\frac{(n1)}{(\sqrt{s}nm)}.$$ I.e., $`E_{kin}=\frac{3}{2}T`$, where $`E_{kin}=(\sqrt{s}nm)`$ is the kinetic energy. It is important to note that the equation(3.4) have unique real rising with $`n`$ and decreasing with $`s`$ solution $`\beta _c(s,n)`$ . The expansion of integral (3.2) near $`\beta _c(s,n)`$ unavoidably gives asymptotic series with zero convergence radii since $`\rho _n(\beta )`$ is the essentially nonlinear function of $`\beta `$. From physical point of view this means that, generally speaking, fluctuations in vicinity of $`\beta _c(s,n)`$ may be arbitrarily high and in this case $`\beta _c(s,n)`$ has not any physical sense. But if fluctuations are small (strictly speaking, they may be arbitrarily high, bu distribution in vicinity of $`\beta _c(s,n)`$ should be Gaussian), then $`\rho _n(\beta )`$ should coincide with partition function of $`n`$ particles and $`\beta _c(s,n)`$ may be interpreted as the inverse temperature. Let us define the conditions when the fluctuations are small . Firstly, we should expand $`\mathrm{ln}\rho _n(\beta +\beta _c)`$ over $`\beta `$: $$\mathrm{ln}\rho _n(\beta +\beta _c)=\mathrm{ln}\rho _n(\beta _c)\sqrt{s}\beta +\frac{1}{2!}\beta ^2\frac{^2}{\beta _c^2}\mathrm{ln}\rho _n(\beta _c)\frac{1}{3!}\beta ^3\frac{^3}{\beta _c^3}\mathrm{ln}\rho _n(\beta _c)+\mathrm{}$$ (3.5) and, secondly, expend the exponent in the integral over, for instance, over $`^3\mathrm{ln}\rho _n(\beta _c)/\beta _c^3`$ neglecting higher decomposition terms in (3.5). In result, $`k`$-th term of the perturbation series $$\rho _{n,k}\left\{\frac{^3\mathrm{ln}\rho _n(\beta _c)/\beta _c^3}{(^2\mathrm{ln}\rho _n(\beta _c)/\beta _c^2)^{3/2}}\right\}^k\mathrm{\Gamma }\left(\frac{3k+1}{2}\right).$$ (3.6) Therefore, one should assume that $$^3\mathrm{ln}\rho _n(\beta _c)/\beta _c^3<<(^2\mathrm{ln}\rho _n(\beta _c)/\beta _c^2)^{3/2}.$$ (3.7) to neglect this term. One of possible solution of this condition is $$^3\mathrm{ln}\rho _n(\beta _c)/\beta _c^30.$$ (3.8) If this condition is hold, then the fluctuations are Gaussian, but arbitrary since theirs value is defined by $`\{^2\mathrm{ln}\rho _n(\beta _c)/\beta _c^2\}^{1/2}`$, see (3.5). Let us consider now (3.8) carefully. We will find computing derivatives that this condition means following approximate equality: $$\frac{\rho _n^{(3)}}{\rho _n}3\frac{\rho _n^{(2)}\rho _n^{(1)}}{\rho _n^2}+2\frac{(\rho _n^{(1)})^3}{\rho _n^3}0,$$ (3.9) where $`\rho _n^{(k)}`$ means the $`k`$-th derivative. For identical particles (see definition (3.3)), $`\rho _n^{(k)}(\beta _c)=n^k(1)^k{\displaystyle \left\{\underset{i=1}{\overset{n}{}}\epsilon (q_i)\frac{d^3q_ie^{\beta \epsilon (q_i)}}{(2\pi )^32\epsilon (q_i)}\right\}|A_n^{ab}|^2}`$ $`=\sigma _{tot}n^k{\displaystyle \left\{\underset{i=1}{\overset{k}{}}\epsilon (q_i)\frac{d^3q_ie^{\beta \epsilon (q_i)}}{(2\pi )^32\epsilon (q_i)}\right\}\overline{f}_k(q_1,q_2,\mathrm{},q_k)},`$ (3.10) where $`\overline{f}_k`$ is the $`(nk)0`$-point inclusive cross section. It coincide with $`k`$-particle distribution function in the $`n`$-particle system. Therefore, l.h.s. of(3.9) is the 3-point correlator $`K_3`$: $$K_3𝑑\omega _3(q)\left(<\underset{i=1}{\overset{3}{}}\epsilon (q_i)>_{\beta _c}3<\underset{i=1}{\overset{2}{}}\epsilon (q_i)>_{\beta _c}<\epsilon (q_3)>_{\beta _c}+2\underset{i=1}{\overset{3}{}}<\epsilon (q_i)>_{\beta _c}\right),$$ (3.11) where the index means averaging with the Boltzmann factor $`\mathrm{exp}\{\beta _c\epsilon (q)\}`$. In result, to have all fluctuations in vicinity of $`\beta _c`$ Gaussian, we should have $`K_m0`$, $`m3`$. But, as follows from (3.7), the set of minimal conditions looks as follows: $$K_m<<K_2,m3.$$ (3.12) If experiment confirms this conditions then, independently from number of particles, the final state may be described by one parameter $`\beta _c`$ with high enough accuracy $`\beta _c`$. Considering $`\beta _c`$ as physical (measurable) quantity, we are forced to assume that both the total energy of the system $`\sqrt{s}=E`$ and conjugate to it variable $`\beta _c`$ may be measured with high accuracy<sup>5</sup><sup>5</sup>5Note, the uncertainty principle $`\mathrm{}`$ did not restrict $`\mathrm{\Delta }E`$ and $`\mathrm{\Delta }\beta `$.. ## 4 Real-time finite temperature generating functionals We would like to show now why and in a what conditions our $`S`$-matrix interpretation of statistics is rightful. In modern formulations, see e.g. the textbook , the temperature is introduced by so called periodic Kubo-Martin-Schwinger (KMS) boundary condition . Namely, in the Feynman-Kac functional integral representation of the partition function $$\mathrm{\Xi }(\beta )=D\phi e^{S_\beta (\phi )}.$$ (4.1) the action $`S_\beta (\phi ;z)`$ is defined on the Matsubara imaginary time contour $`C_M`$: $`(t_i,t_ii\beta )`$, but fields should obey KMS boundary condition: $$\phi (t_i)=\phi (t_ii\beta ).$$ (4.2) This is natural consequence of definition: $`\mathrm{\Xi }(\beta )=\mathrm{Sp}e^{\beta 𝐇}`$. It was offered to deform Matsubara contour in a following way: $$C_MC_{SK}:(t_i,t_f)+(t_f,t_i+i\beta ),$$ (4.3) where $`C_{SK}`$ is the Mills time contour and $`t_f>t_i`$ belongs to real axis . Including the real-time parts we obtain a possibility to describe time evolution of the system But this attempt was not successful. First of all, we have not an evident interpretation $`t_i`$ and $`t_f`$ . Secondly, in spite of real-time parts, this formulation unable to describe the time evolution . ### 4.1 Equilibrium media It was shown above that if $`\sigma _n`$ is defined by (3.1) then one may introduce $`\rho _n`$ using definition (3.3). The Fourier transform (3.2) connects $`\sigma _n`$ $`\rho _n`$. On other hand, $`\rho _n`$ reminds the partition function. To find complete analogy with statistical physics we should consider transition $`mn`$ particles with amplitude $`A_{nm}=<out;n|in;m>`$. Summation over $`n`$ and $`m`$ is assumed. The corresponding $`\delta `$-function of energy-momentum conservation law should be written in the form: $$\delta (\underset{i=1}{\overset{n}{}}q_i\underset{i=1}{\overset{m}{}}p_i)=d^4P\delta (P\underset{i=1}{\overset{n}{}}q_i)\delta (P\underset{i=1}{\overset{m}{}}p_i),P=(E,\stackrel{}{P}).$$ (4.4) This will lead to necessity introduce independently the temperature of initial ($`1/\beta _i`$) and final ($`1/\beta _f`$) states. In particle physics we can consider the final state temperature only. In result we get to the Fourier-Mellin transform $`\rho (\beta ,z)=\rho (\beta _i,z_i;\beta _f,z_f)`$. Direct calculations give important factorized form: $$\rho (\beta ,z)=e^{\widehat{N}(\beta ,z;\varphi )}\rho _0(\varphi ),$$ where the operator $`\widehat{N}(\beta ,z;\varphi )={\displaystyle }dxdx^{}(\widehat{\varphi }_+(x)D_+(xx^{},\beta _f,z_f)\widehat{\varphi }_{}(x^{})`$ $`\widehat{\varphi }_{}(x)D_+(xx^{},\beta _i,z_i)\widehat{\varphi }_+(x^{})),\widehat{\varphi }={\displaystyle \frac{\delta }{\delta \varphi }},`$ (4.5) acts on the functional: $$\rho _0(\varphi _\pm )=D\mathrm{\Phi }_+D\mathrm{\Phi }_{}e^{iS(\mathrm{\Phi }_+)iS(\mathrm{\Phi }_{})iV(\mathrm{\Phi }_++\varphi _+)+iV(\mathrm{\Phi }_{}+\varphi _{})}.$$ (4.6) At the very end of calculations one should take auxiliary variables $`\varphi _\pm `$ equal to zero. Here $`D_+`$ $`D_+`$ are the frequency correlation functions: $$D_\pm (xx^{},\beta )=i𝑑\omega (q)e^{\pm iq(xx^{}+i\beta )}z(q)$$ They obey the equations: $$(^2+m^2)_xG_+=(^2+m^2)_xG_+=0.$$ So, all ‘thermodynamical’ information contained in the operator $`\widehat{N}(\beta ,z;\varphi )`$, but interactions are described by $`\rho _0(\varphi )`$. One can say that the operator $`\widehat{N}`$ (adiabatically) maps the interacting filed system on the observable states. This important property allows consider only ‘mechanical’ processes and exclude from consideration the ‘thermal’ ones. Calculating $`\rho _0(\varphi )`$perturbatively one can find: $$\rho (\beta ,z)=e^{iV(i\widehat{j}_+)+iV(i\widehat{j}_{})}e^{\frac{i}{2}{\scriptscriptstyle 𝑑x𝑑x^{}j_a(x)D_{ab}(xx^{},\beta ,z)j_b(x^{})}},$$ (4.7) where $`D_{++}`$ is the Feynman (causal) Green function and $$D_{}=(D_{++})^{}$$ is the anticausal one and, as usual, $`\widehat{j}=\delta /\delta j`$. At the very end one should take $`j=0`$. Let us assume now that our system is a subsystem of bigger system. This would lead to transformation of Boltzmann factor $`\mathrm{exp}\{\beta \epsilon \}`$ on corresponding statistics occupation number $`\overline{N}(\beta \epsilon )`$. This means that our interacting fields system is surrounded by black body radiation. This is mechanical model of the thermostat (heat bath of thermodynamics). In result the matrix $`D_{ab}`$ takes form (we put for simplicity $`z_i=z_f=1`$): $`iG(q;\beta )=\left(\begin{array}{cc}\frac{i}{q^2m^2+iϵ}& 0\\ 0& \frac{i}{q^2m^2iϵ}\end{array}\right)+`$ $`+2\pi \delta (q^2m^2)\left(\begin{array}{cc}\stackrel{~}{n}(\frac{\beta _f+\beta _i}{2}|q_0|)& \stackrel{~}{n}(\beta _i|q_0|)a_+(\beta _i)\\ \stackrel{~}{n}(\beta _f|q_0|)a_{}(\beta _f)& \stackrel{~}{n}(\frac{\beta _f+\beta _i}{2}|q_0|)\end{array}\right)`$ (4.8) where $$a_\pm (\beta )=e^{\frac{\beta }{2}(|q_0|\pm q_0)}.$$ Following Green functions: $$D_{ab}(xx^{},\beta )=\frac{d^4q}{(2\pi )^4}e^{iq(xx^{})}G_{ab}(q,\beta )$$ was introduced and the occupation number $$n_{++}(q_0)=n_{}(q_0)=\left\{e^{|q_0|(\beta _f+\beta _i)/2}1\right\}^1\stackrel{~}{n}(|q_0|\frac{\beta _i+\beta _f}{2}).$$ (4.9) and $$n_+(q_0)==\mathrm{\Theta }(q_0)(1+\stackrel{~}{n}(q_0\beta _f))+\mathrm{\Theta }(q_0)\stackrel{~}{n}(q_0\beta _i),$$ (4.10) $$n_+(q_0)=\mathrm{\Theta }(q_0)\stackrel{~}{n}(q_0\beta _i)+\mathrm{\Theta }(q_0)(1+\stackrel{~}{n}(q_0\beta _f)).$$ (4.11) Assuming that $`\beta _i=\beta _f=\beta _c`$ it is easy to find: $$G_+(tt^{})=G_+(tt^{}i\beta ),G_+(tt^{})=G_+(tt^{}+i\beta ),$$ (4.12) i.e. our Green function obey KMS boundary condition. So, representation (4.7) with Green functions (4.8) coincide identically with (4.1) calculated perturbatively, see also . ### 4.2 Nonequilibrium media Our attempt introduce the temperature as the quantitative characteristic of $`whole`$ system based on assumption that mean value of correlators is small. We can ‘localize’ this condition assuming that this rough description may be extended only on subdomains of the system. For definiteness the subdomains may be marked by space-time coordinate $`r`$. It should be underlined that we divide on the subdomains not the system under consideration but the device, where external particles are measured. Noting that external flow consist from noninteracting particles (including the flow of black body radiation) the division on subdomains can not influence on the fields interaction. In result we introduce the ‘local’ temperature $`1/\beta (r)`$ for $`r`$-th group of interacting particles assuming that fluctuations in vicinity of $`\beta (r)`$ are Gaussian. This means that the mean value of correlation in the group is small, but the correlation between groups may be high. Nevertheless last one is not important since the external particles are on the mass shell. At the same time dimension of group may be arbitrary, but large then some $`r_0`$ to have possibility to introduce the temperature as the collective variable. We can distinguish following scales. Let $`L_q`$ be the characteristic 4-scale of quantum fluctuations, $`L_s`$ be the scale thermodynamical fluctuations and $`L`$ be the scale of subdomain. It is natural to assume that $`L_s>>L>>L_q`$. Corresponding generating $`functional`$ has the form: $$\rho _{cp}(\alpha _1,\alpha _2)=e^{\widehat{N}(\varphi _a^{}\varphi _b)}\rho _0(\varphi _\pm ).$$ One may note that the ‘localization’ gives influence on the operator only: $$\widehat{N}(\varphi _a^{}\varphi _b)=𝑑Y𝑑y\widehat{\varphi }_a(Y+y/2)\stackrel{~}{n}_{ab}(Y,y)\widehat{\varphi }_b(Yy/2),$$ The occupation numbers $`n_{ab}(Y,q)`$ have same form, $`\beta \beta (Y)`$ and $$\stackrel{~}{n}_{ij}(Y,y)=𝑑\omega (q)e^{iqy}n_{ij}(Y,q)$$ We find calculating $`\rho _0`$ perturbatively that: $`\rho _{cp}(\beta )=\mathrm{exp}\{iV(i\widehat{j}_+)+iV(i\widehat{j}_{})\}\times `$ $`\mathrm{exp}\{i{\displaystyle }dYdy[j_a(Y+y/2)G_{ab}(y,(\beta (Y))j_b(Yy/2)\}`$ (4.13) where the matrix Green function $`G(q,(\beta (Y)))`$ was defined in (4.8). ## 5 Conclusion One more detail. Our consideration has show the uniqueness of Bogolyubov’s solution of the nonequilibrium thermodynamics problem. Indeed, without vanishing of correlations perturbation series in the $`\beta _c`$ vicinity, being asymptotic, is divergent. We would like to stress in conclusion that Bogolyubov’s creative works naturally unite particle and statistical physics. In result, using Bogolyubov’s mathematical basis, we have the united scientific space in which both branches of physics, thermodynamics and quantum field theory, supplement each other. Acknowledgement We would like to thank V.Kadyshevsky, V.A.Matveev, A.N.Tavkhelidze for interest to described approach and E.Kuraev for fruitful discussions.
warning/0003/hep-ph0003197.html
ar5iv
text
# 1 Introduction ## 1 Introduction There is accumulating evidence that the expansion rate of the universe is greater now than in the past . The data, taken alone or in conjunction with constraints from large scale structure , are compatible (in a flat universe) with a contribution $`\mathrm{\Omega }_\mathrm{\Lambda }0.7`$ from a cosmological constant $`\mathrm{\Lambda }`$ or its equivalent. Cluster abundance estimates for the matter fraction $`\mathrm{\Omega }_{matter}`$ of the critical density, combined with an analysis of CMB observations in the Doppler peak region (which support the inflation prediction $`\mathrm{\Omega }_{tot}=1`$) are also compatible with such a contribution to the “dark energy”. The resulting energy density is given by $$\rho _\mathrm{\Lambda }=(2.2\times 10^3\mathrm{eV})^4(\mathrm{\Omega }_\mathrm{\Lambda }/0.7)(h/0.65)^2,$$ (1) where $`h=`$ present Hubble constant $`H_0`$ in units of 100 km/sec/Mpc. The introduction of a non-zero value for $`\rho _\mathrm{\Lambda }`$ or its equivalent presents an important challenge to particle physicists and cosmologists. Even if one concedes ignorance and simply accepts as premise that the universe is relaxing to a state with $`\rho _\mathrm{\Lambda }=0`$ , there still remains the perplexing question of the origin of a mass scale $`\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}10^3`$ eV which makes $`\rho _\mathrm{\Lambda }`$ relevant in the present era. Discussion in recent years has centered on models in which $`\mathrm{\Lambda }`$ becomes time-dependent, originating in the energy density of a scalar field $`\varphi `$ evolving in a potential $`V(\varphi ).`$ One candidate for $`\varphi `$ is a pseudo-Nambu-Goldstone boson (axion) in a harmonic potential associated with the breaking of a $`U(1)`$ symmetry to $`Z_N`$ . The cosmological consequences of this scenario depend on both the normalization of the potential $`M^4`$ and the scale $`f`$ at which the $`U(1)`$ symmetry is realized in the Goldstone mode, and there have been studies where the parameters make the model relevant to late-time cosmology, including recent applications to the SNeIa results . The required scale $`M10^3\text{eV}`$ can be associated with a neutrino mass or with the confining scale of a hidden gauge theory , $`\varphi `$ being the axion. In quintessence models , $`\varphi `$ evolves so that the late-time behavior of the dark energy density $`\rho _\varphi `$ is largely independent of initial conditions. What remains to be tuned by hand is the parameter in the potential which allows for a positive acceleration of the scale parameter during the SNeIa era relevant to the observations of Ref. . The origin of the field $`\varphi ,`$ the form of its potential, and its relation to other physics remain to be explained . In this paper, I would like to propose that the dark energy $`\rho _\mathrm{\Lambda }`$ is not evolving, but is the false vacuum energy associated with an incomplete chiral phase transition in a hidden $`SU(2)^{}`$ gauge theory with strong scale $`\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}.`$ It will be seen that in the context of modern D-brane physics, the scale $`\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}`$ for the vacuum energy can be accommodated in a natural manner in a supersymmetric GUT theory. Instead, the central problem will be to explain how a low temperature $`T_{hidden}<T_{\mathrm{CMB}}\frac{1}{10}\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}`$ can be sustained for the quark-gluon phase of the (supercooled) plasma of the hidden sector. Such a high degree of supercooling can potentially be tested in lattice simulations. In the present work, it will strongly constrain the effective field theory. The model is described in the next section, and some possible advantages as an alternative to the scalar field scenario are mentioned in the concluding section. ## 2 The Model I will consider a hidden unbroken $`SU(2)^{}`$ Yang-Mills theory whose low energy matter content is $`N_f`$Dirac fermions (2$`N_f`$Weyl fermions) with vector-like coupling to the gauge field. (For now, hidden sector quantities will be denoted by a prime.) Except for gravitational interactions, this $`SU(2)^{}`$ is completely decoupled from all standard model fields. The choice of $`SU(2)^{}`$ is doubly motivated: (1) the running of the gauge coupling is slow, so that the scale $`\mathrm{\Lambda }_{SU(2)^{}}`$ can be pushed to values approximating $`\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}`$ (2) there will be fewer massless degrees of freedom to perturb the successful scenario of Big Bang Nucleosynthesis (BBN), a critical requirement of the model. The hidden nature of the $`SU(2)^{}`$ will be discussed when the coupling requirement at GUT energies is obtained. Since the theory will be considered to have evolved from GUT energies, the supersymmetric version becomes relevant. The matter content then consists solely of 2$`N_f`$chiral $`SU(2)^{}`$ doublet superfields $`Q_i(i=1\mathrm{}2N_f).`$ In order to preserve the low energy chiral phase transition, $`SU(2)^{}`$ singlet, flavor antisymmetric mass terms $`Q_iQ_jQ_jQ_i`$ must be prohibited by a discrete symmetry $`(R`$-invariance or $`Z_N,N>2`$ symmetry). Next, the 1-loop RG equation relates the gauge coupling at GUT $`\alpha _{\mathrm{GUT}}^{}`$ and the strong coupling scale $`\mathrm{\Lambda }_{SU(2)^{}}`$: $`{\displaystyle \frac{2\pi }{\alpha _{\mathrm{GUT}}^{}}}`$ $`=`$ $`b_2^{\mathrm{SUSY}}\mathrm{ln}\left({\displaystyle \frac{M_{\mathrm{GUT}}}{1\mathrm{TeV}}}\right)+b_2^{\mathrm{non}\mathrm{SUSY}}\mathrm{ln}\left({\displaystyle \frac{1\mathrm{TeV}}{\mathrm{\Lambda }_{SU(2)^{}}}}\right)`$ $`b_2^{\mathrm{SUSY}}`$ $`=`$ $`6N_f`$ $`b_2^{\mathrm{non}\mathrm{SUSY}}`$ $`=`$ $`\frac{22}{3}\frac{2}{3}N_f,`$ (2) so that for $`M_{\mathrm{GUT}}=2\times 10^{16}\mathrm{GeV},`$ $`\mathrm{\Lambda }_{SU(2)^{}}=\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}=2.2\times 10^3\mathrm{eV}`$ one obtains $$\alpha _{\mathrm{GUT}}^{}{}_{}{}^{1}=68.68.46N_f.$$ (3) The major premise of the model is the existence of a false vacuum at present. A first order phase transition driven by fluctuations of chiral condensates is strongly indicated by theoretical arguments , most cleanly for $`N_f4.`$ From Eq. (3), it would seem that unification with standard model couplings ($`\alpha _{\mathrm{GUT}}\frac{1}{25}`$) can be achieved with a choice $`N_f=5;`$ however, this would elevate the number of Goldstone degrees of freedom below the critical temperature $`(=2N_f^2N_f1)`$ to a value larger than the effective number of quark-gluon degrees of freedom $`(=7N_f+6),`$ presumably vitiating the first order phase transition. Thus I choose $`N_f=4,`$ and Eq. (3) gives $$\alpha _{\mathrm{GUT}}^{}\frac{1}{35}\alpha _{\mathrm{GUT}}.$$ (4) Such a disparity in GUT-scale gauge couplings is not difficult to accommodate in current formulations of string theory, with gauge fields residing in open strings tied to D-branes. If, for example, the standard model gauge group lives on a 5-brane and the hidden $`SU(2)^{}`$ on another 5-brane (orthogonal with respect to the compactified 2-tori) , the ratio of the gauge couplings would be inversely proportional to the volumes of the 2-tori: $$\frac{\alpha _{\mathrm{GUT}}^{}}{\alpha _{\mathrm{GUT}}}=\frac{v_2}{v_2^{}}.$$ (5) Thus, a 20% difference in toroidal moduli could account for the disparity in the $`\alpha `$’s. ### Temperature Constraints Having established a model, one can quickly ascertain the constraints which follow from BBN. The hidden sector energy density $`\rho ^{}`$ of the $`SU(2)^{}`$ gauge fields and 2$`N_f`$ Weyl doublets, relative to a single species of left-handed neutrino is given by $$\frac{\rho ^{}}{\rho _{\nu _e}}|_{BBN}=\left(\frac{7N_f+6}{(7/4)}\right)\left(\frac{T^{}}{T}\right)^4|_{BBN}.$$ (6) Requiring this ratio to be $`0.3`$ implies (for $`N_f=4)`$ $$\frac{T^{}}{T}|_{BBN}0.353.$$ (7) Much of this can be accounted for through reheat processes in the visible sector. Assuming no reheat for the hidden sector for energies below the electroweak scale, one finds $`{\displaystyle \frac{T^{}}{T}}|_{BBN}`$ $`=`$ $`\left({\displaystyle \frac{g^{}(BBN)}{g^{}(>EW)}}\right)^{1/3}{\displaystyle \frac{T^{}}{T}}|_{>EW}`$ (8) $`=`$ $`\left({\displaystyle \frac{10.75}{106.75}}\right)^{1/3}{\displaystyle \frac{T^{}}{T}}|_{>EW}=0.465{\displaystyle \frac{T^{}}{T}}|_{>EW},`$ (9) where $`>EW`$ denotes temperatures above the electroweak scale. In order to comply with the BBN requirement (Eq. 6) an additional suppression of $`T^{}/T`$ above the electroweak scale by a factor of 0.76 is required. Here one can invoke an asymmetric post-inflation reheat into visible and hidden sector quanta . In the slow reheat scenario , $`T_{reheat}`$ is proportional to the coupling of the inflaton to the quanta , so that a ratio of couplings of the same order as the ratio of the $`\alpha `$’s (4) could provide the desired additional suppression. This reheat asymmetry could have the same origin as the $`\alpha `$ asymmetry, if the inflaton originates in the modular sector. Asymmetric reheating also obtains in the case of parametric resonance because of the asymmetric coupling . Because of $`e^+e^{}`$ annihilation, the present $`T^{}`$ is further depressed relative to the present CMB temperature by the same $`(4/11)^{1/3}`$ factor as with neutrinos. Thus, together with (7), one obtains $`(T_{\mathrm{now}}^{}/T_{\mathrm{CMB}})0.251.`$ Since $`T_{\mathrm{CMB}}=2.35\times 10^4\mathrm{eV}=0.11\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4},`$ one finds $$T_{\mathrm{now}}^{}/\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}0.028.$$ (10) This suggests a great deal of supercooling, which needs to be accommodated. The calculations in this work will require the ratio $`T/T_c`$ ($`T_c`$ is the critical temperature<sup>1</sup><sup>1</sup>1From here on, all temperatures will be understood to be hidden sector temperatures, and the primes will be omitted.) which in turn requires knowledge of the ratio $`\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}/T_c.`$ This will be calculable in the effective field theory to be discussed. ### Supercooling In the standard formulation of first order phase transitions via critical bubble formation the condition for failure to complete a phase transition in the expanding universe at (hidden) temperature $`T`$ is $$(T/H(T))^4e^{F_c/T}<1,$$ (11) where $`F_c`$ is the free energy of a critical bubble, and $`H(T)`$ is the Hubble constant at temperature $`T.`$ With $`H_02.2\times 10^{33}h`$ eV, $`T=0.28T_{\mathrm{CMB}},`$ one obtains the condition for failure to nucleate $$F_c/T>260.$$ (12) In the thin wall approximation, the bubble has a well-defined surface tension $`\sigma ,`$ and the picture is consistent only for small supercooling below $`T_c.`$ The bubble action is given by $$\frac{F_c}{T}=\frac{16\pi }{3}\frac{\sigma ^3}{L^2\eta ^2T_c},$$ (13) where $`L`$ is the latent heat and $`\eta =(T_cT)/T_c.`$ Thus, a failure to nucleate via thin-walled bubbles requires a large surface tension, $`\sigma /T_c^3\stackrel{>}{}1.`$ In lattice studies of quenched QCD , the interface tension between confined and deconfined phases is small: $`\sigma /T_c^30.1.`$ However, a simple calculation based on the MIT bag model suggests that the picture can change drastically in the presence of chiral condensates: in that case, $$\sigma =\frac{1}{4}\underset{i=1}{\overset{N_f}{}}\overline{q}_iq_i.$$ (14) which for QCD $`(T_c150\text{MeV},\overline{q}_iq_i(240\text{MeV})^3`$ per flavor) would give a large surface tension, $`\sigma 4T_c^3,`$ possibly invalidating the thin-wall approximation.<sup>2</sup><sup>2</sup>2This would give an average distance between nucleation sites of about 10 m , perhaps marginally affecting primordial light element abundances . A full analysis of the unquenched QCD situation is probably best carried in the framework of a mean field theory . I will proceed in the context of such a theory to see what constraints are imposed on the $`SU(2)^{}`$ model in order to attain the desired metastability (Eq. (12)) until the present era. ## 3 Linear Sigma Model for $`SU(2)^{}`$ with $`N_f`$Flavors. The symmetry breaking pattern of color $`SU(2)^{}`$ with $`N_f`$flavors has long been known , and a linear sigma model for this case has recently been examined . Such a model will serve conveniently to study the chiral phase transition. The meson and diquark baryon fields are contained in the $`2N_f\times 2N_f`$ antisymmetric matrix $`\mathrm{\Phi }_{ij}=\mathrm{\Phi }_{ji},`$ with chiral symmetry breaking occurring in $`Sp(2N_f)`$ direction compatible with the Vafa-Witten theorem $$\mathrm{\Phi }_0=\frac{\varphi _0}{\sqrt{2(2N_f)}}\left(\begin{array}{ccc}\mathrm{𝟎}& \mathrm{𝟏}& \\ \mathrm{𝟏}& \mathrm{𝟎}& \end{array}\right)$$ (15) where $`\mathrm{𝟎}`$ and $`\mathrm{𝟏}`$ are $`N_f\times N_f`$ matrices. The lagrangian is $$=\text{Tr}^\mu \mathrm{\Phi }_\mu \mathrm{\Phi }m^2\text{Tr}\mathrm{\Phi }^{}\mathrm{\Phi }\lambda _1\left(\text{Tr}\mathrm{\Phi }^{}\mathrm{\Phi }\right)^2\lambda _2\text{Tr}\mathrm{\Phi }^{}\mathrm{\Phi }\mathrm{\Phi }^{}\mathrm{\Phi }.$$ (16) I have omitted a term $`\text{Pf}(\mathrm{\Phi })+\text{Pf}(\mathrm{\Phi }^{})`$ arising from the axial anomaly. Since I will be working with $`N_f=4,`$ this additional operator quartic in the fields will not qualitatively change the discussion which follows. Stability in all field directions requires $`\lambda _20,\lambda _1+\lambda _2/2N_f0.`$ Specializing now to the field $`\varphi `$ in the direction of the vev, one obtains $`_0`$ $`=`$ $`\frac{1}{2}(_\mu \varphi )^2V(\varphi ),`$ $`V(\varphi )`$ $`=`$ $`\frac{1}{2}m^2\varphi ^2+\frac{1}{4}\overline{\lambda }\varphi ^4,`$ $`\overline{\lambda }`$ $`=`$ $`\lambda _1+\overline{\lambda }_2,\overline{\lambda }_2=\lambda _2/2N_f.`$ (17) With finite temperature corrections (restricted for simplicity to the $`m^2`$ term) and the introduction of a running quartic coupling, one obtains the effective potential $$V(\varphi ,T)=\frac{1}{2}m^2(T)\varphi ^2+\frac{1}{4}\overline{\lambda }(t)\varphi ^4,$$ (18) where $`t=\mathrm{ln}(\varphi /\varphi _0)`$, and $`m^2`$ has the standard $`T`$-corrected form $$m^2(T)=A(T^2T_0^2).$$ (19) In the model described, $`A`$ can be calculated, and I find at one loop $$A=\frac{1}{12}\left[(2N_f(2N_f1)+2)\overline{\lambda }+(6N_f(2N_f1)2)\overline{\lambda }_2\right].$$ (20) The $`ϵ`$ expansion analysis of the model described by (16) shows that it allows a first order phase transition through a Coleman-Weinberg mechanism at $`T=T_0,`$ when $`m^2(T)=0`$ . However, it will shortly be apparent that the large supercooling will require that $`T_0^2T_c^2.`$ This (approximate) conformal invariance at tree level in the chiral lagrangian (to be discussed below) in turn suggests that chiral symmetry breaking at zero temperature in this model also proceeds through radiative corrections (Coleman-Weinberg) . Thus, to lowest order (see Eq. (29) below) $$\overline{\lambda }(t)=\lambda (14t),\lambda \overline{\lambda }(0).$$ (21) where $`t=0`$ is defined by the minimum of the second term in (18). The vacuum at $`\varphi =0`$ described by the potential (18) becomes metastable at a temperature $`T_c`$ determined by requiring simultaneously $$V^{}(\varphi _+)=V(\varphi _+)=0,$$ (22) at some field value $`\varphi _+.`$ A short algebraic exercise with Eqs. (18), (19), (21), and (22) determines $`T_c:`$ $$m(T_c)=\sqrt{A(T_c^2T_0^2)}=\sqrt{\lambda }\varphi _0e^{1/4}.$$ (23) ### Suppression of $`m^2`$. For $`T_0<T<T_c`$ there is a barrier between the false and true vacua. But the extreme supercooling requirement indicated in Eq. (10) will be seen to impose a large hierarchy, $`T_0T_c`$ (it will turn out that $`T_00.124T_c.)`$ There is no obvious argument to justify this hierarchy — it is simply required in this model for compatibility with the supercooling requirement. Nevertheless, two comments may be made: (1) A quantitative comparison with the linear $`SU(3)\times SU(3)`$ sigma model (whose dynamics differs through the presence of the cubic term) is perhaps instructive. In that case, the coefficient $`A`$ is obtained as a sum over the massive mesons , $`A=\frac{1}{12}_iM_i^2/v^2,`$ where $`vf_\pi .`$ With $`M_i1`$ GeV, one estimates $`A83.`$ Since $`T_0=\sqrt{\mu ^2/A},`$ where $`\mu ^20.15\text{GeV}^2`$ is the temperature-independent (negative) mass parameter, we find $`T_043`$ MeV $`0.24T_c.`$ This may provide a normative expectation for $`T_0/T_c.`$ (2) In $`SU(N)`$ gauge theories with $`N_f`$ massless fermion flavors, a continuous transition to an approximately conformal phase (which would imply $`m^2=0)`$ is suggested at some value of $`N_f/N<11/2\text{[39]}.`$ Some analytic studies indicate $`N_f/N4`$ as a critical value, while a QCD lattice study show hints of a suppression of the chiral condensate (expected in the transition to the conformal phase ) for a smaller value, $`N_f/N=\frac{4}{3}`$ ($`SU(3)`$ with 4 flavors). Perhaps the ratio in the present case $`(N_f/N=2)`$ is sufficiently large to significantly suppress the zero temperature $`m^2`$ term in the effective lagrangian — a lattice study of chiral symmetry breaking in $`SU(2)`$ with $`N_f=4`$ could in principle shed light on this question. At any rate, at this juncture I accept the hierarchy $`T_0T_c,`$ and simplify matters even more by setting $$T_0=0$$ (24) in Eq. (19). In such a model, with two coupling constants, the chiral invariance is broken at $`T=0`$ in the Coleman-Weinberg manner . For $`T0,`$ the transition becomes first order, as described in the previous section. ### Critical Bubbles. For $`T0`$ it will prove convenient to rescale $`\varphi =m(T)\varphi ^{}/2\sqrt{\lambda },`$ so that using Eqs. (18), (21) and (23) one may write $`V`$ $``$ $`{\displaystyle \frac{m^4(T)}{4\lambda }}\overline{V}`$ $`\overline{V}`$ $`=`$ $`\frac{1}{2}\varphi _{}^{}{}_{}{}^{2}+\frac{1}{4}\varphi _{}^{}{}_{}{}^{4}\left[\mathrm{ln}((m(T)/m(T_c))\varphi ^{})/2\frac{1}{2}\right]`$ (25) $`=`$ $`\frac{1}{2}\varphi _{}^{}{}_{}{}^{2}+\frac{1}{4}\varphi _{}^{}{}_{}{}^{4}\left[\mathrm{ln}((T/T_c)\varphi ^{}/2)\frac{1}{2}\right]`$ for $`T_0=0.`$ The O(3) symmetric free energy for a critical bubble formed at temperature $`T`$ is given by $`F_c`$ $`=`$ $`4\pi {\displaystyle \frac{m(T)}{4\lambda }}{\displaystyle _0^{\mathrm{}}}𝑑r^{}r_{}^{}{}_{}{}^{2}\left[\frac{1}{2}\left(d\varphi ^{}/dr^{}\right)^2+\overline{V}(\varphi ^{},T/T_c)\right]`$ (26) $``$ $`{\displaystyle \frac{m(T)}{4\lambda }}f(T/T_c)`$ where $`r^{}=m(T)r.`$ The field $`\varphi ^{}`$ is the solution to $$\frac{d^2\varphi ^{}}{dr_{}^{}{}_{}{}^{2}}+\frac{2}{r^{}}\frac{d\varphi ^{}}{dr^{}}=\frac{\overline{V}}{\varphi ^{}}.$$ (27) subject to $`d\varphi ^{}/dr^{}|_{r^{}=0}=0,\varphi ^{}(\mathrm{})=0.`$ With the help of Eqs. (19), (20) and (24) the bubble action can then be calculated more explicitly in terms of the the quantity $`f(T/T_c).`$ For $`N_f=4,`$ I obtain $`F_c/T`$ $`=`$ $`\sqrt{A}/(4\lambda )f(T/T_c)`$ (28) $`=`$ $`\sqrt{\frac{29}{6}\overline{\lambda }+\frac{83}{6}\overline{\lambda }_2}/(4\lambda )f(T/T_c).`$ At this point, I implement the condition of radiative symmetry breaking (at $`T=0`$) in the two-parameter $`(\overline{\lambda },\overline{\lambda }_2)`$ space. This imposes the condition $$4\overline{\lambda }(0)=4\lambda =\beta _{\overline{\lambda }}(0).$$ (29) From the one-loop RG equations , for $`\overline{\lambda }_2(0)\left|\overline{\lambda }(0)\right|,`$ but $`\overline{\lambda }_2(0)`$ still perturbative (this will be justified a posteriori), Eq. (29) gives $$\overline{\lambda }_2(0)=\left(4\pi /\sqrt{4N_f^22N_f2}\right)\sqrt{\overline{\lambda }(0)}=(4\pi /\sqrt{54})\sqrt{\lambda },$$ (30) and hence in the same approximation $$A\frac{83}{6}\frac{4\pi }{\sqrt{54}}\sqrt{\lambda }.$$ (31) Thus the bubble action can be written entirely in terms of the coupling constant $`\lambda =\overline{\lambda }(0):`$ $`F_c/T`$ $`=`$ $`\sqrt{\frac{4\pi }{\sqrt{54}}\frac{83}{6}\sqrt{\lambda }}/(4\lambda )f(T/T_c)`$ (32) $``$ $`1.22f(T/T_c)\lambda ^{3/4}.`$ We now require the ratio $`T/T_c.`$ From Eq. (10), one needs to relate the vacuum energy $`\rho _\mathrm{\Lambda }`$ to $`T_c.`$ At $`T=0,m(T)=0,`$ Eqs. (18), (21), (23), (24) and (31) give $`\rho _\mathrm{\Lambda }`$ $`=`$ $`V(0)V(\varphi _0)`$ (33) $`=`$ $`\frac{1}{4}\lambda \varphi _0^4`$ $`=`$ $`\frac{1}{4}eA^2T_c^4/\lambda `$ $`=`$ $`(4.4T_c)^4.`$ Combining this with Eq. (10), we have the supercooling requirement $$T/T_c0.124.$$ (34) The bubble action may now be evaluated numerically, and I find $`f(0.124)=6.61.`$ Since $`f(T/T_c)`$ is a uniformly decreasing function of $`T/T_c,`$ I obtain (using (12)) the condition for no nucleation until the present era $$260F_c/T(1.22)(6.61)\lambda ^{3/4},$$ (35) giving a bound on $`\lambda ,`$ $$\lambda 0.010,\text{or}0.010\overline{\lambda }(0)0.$$ (36) ### Fine Tuning? Does the bound (36) represent a substantial fine tuning? In an attempt to answer this question, I have presented in Figure 1 the renormalization group flow in the $`\overline{\lambda }\overline{\lambda }_2`$ plane over less than or equal to a decade $`\left[(t\mathrm{ln}(10))(t=0)\right]`$ for those trajectories which begin in the stability region and satisfy the requirement (36). It is seen that (1) a reasonable piece of the coupling constant phase space is available (2) the couplings are perturbative but not particularly small over much of the RG flow and (3) over some of the phase space $`\overline{\lambda }`$ reaches values of the order of the electroweak Higgs coupling. (In the same normalization, $`\lambda _H=0.08(m_H/100\text{GeV})^2.)`$ A summary of results and concluding remarks follows. ## 4 Discussion and Conclusions $`(1)`$A model has been presented which can generate a cosmological constant of magnitude to account for $`\mathrm{\Omega }_\mathrm{\Lambda }0.7`$ during the present era. The present quasi-deSitter phase is driven by the false vacuum energy associated with the supercooled phase of an incomplete chiral phase transition in a hidden gauge theory. The very small energy scale $`\rho _\mathrm{\Lambda }^{\mathrm{\hspace{0.33em}1}/4}2\times 10^3`$ eV for the vacuum energy appears as the strong interaction scale for a hidden $`SU(2)^{}`$ whose coupling runs from GUT energies, with coupling at GUT not quite unifying with the standard model couplings. Having this $`SU(2)^{}`$ and the standard model fields reside on different branes presents a simple solution to this disparity. $`(2)`$An unchanging vacuum energy has some advantages over evolving primordial scalar fields as an origin of the present near-deSitter phase: the problem of protecting the tiny curvature of the potential is circumvented, as is the necessity (in quintessence models) to control the contribution of dark energy during nucleosynthesis . $`(3)`$BBN considerations imply large supercooling in the metastable quark-gluon plasma of the present phase. Although present lattice simulations indicate only small supercooling in the quenched approximation of Yang-Mills theory, theoretical considerations indicate that large supercooling is possible in the presence of chiral condensates. A linear sigma model with two quartic couplings was analyzed in which all symmetry breaking takes place via dimensional transmutation. In this model the existence of substantial supercooling does not require fine tuning in the coupling constant space. The existence of a large interface energy in an $`SU(2)`$ theory with $`N_f=4`$ dynamical quarks would provide incisive support for this model. $`(4)`$A pivotal requirement in this scenario is the near-scale invariance of the chiral lagrangian. This was briefly discussed in the text in terms of the ratio $`N_f/N(=2`$ in the present work). A hint of the continuous approach to a conformal phase may be suggested in the observed weakening of the chiral phase transition in a lattice study of four flavor QCD with $`N_f/N`$ as small as 4/3. $`(5)`$The bound $`\lambda 0.010`$ (Eq. (36)) does not change drastically with a tighter bound on the extra effective number of neutrinos. For example, with a requirement $`\mathrm{\Delta }N_{eff}0.1,`$ I find instead of (36) the bound $`\lambda 0.0086.`$ This would also involve a small amount of extra reheating in the visible sector between the GUT and electroweak scales. $`(5)`$Although the phase transition discussed in this paper is long overdue, it may not be catastrophic when it occurs. The nucleation occurs via very thick-walled bubbles, so that the drastic shock-wave scenario depicted in the thin-walled case is perhaps not inevitable. ### Acknowledgement I would like to thank Gary Steigman for comments and for drawing attention to an arithmetic error in Eq. (9) in the original version of the manuscript. This research was supported in part by the National Science Foundation through Grant No. PHY-9722044.
warning/0003/quant-ph0003035.html
ar5iv
text
# One Complexity Theorist’s View of Quantum ComputingBased on a talk presented at the Second Workshop on Algorithms in Quantum Information Processing at DePaul University, Chicago, January 21, 1999. Realaudio of the talk is available at http://www.cs.depaul.edu/aqip99. ## 1 Introduction When one starts looking at quantum computation an obvious question comes to mind: Can quantum computers be built? By this I mean can one create a large reliable quantum computer—perhaps to factor numbers using Shor’s algorithm \[Sho97\] faster than any classical computer can factor. There has been some success creating in creating quantum machines with a tiny number of bits, we have many physical and engineering issues to overcome before a large scale quantum machine can be realized. As a computational complexity theorist, I would like to consider a different and perhaps more important question: Are quantum computers useful? In other words, even if we can build reasonable quantum machines, will they actually help us solve important computational problems not attainable by traditional machines. Realize that the jury is still out on this issue. Grover’s algorithm \[Gro96\] gives only a quadratic speed-up, which means that a quantum machine will have to approach speeds of a traditional computer (on the order of a trillion operations per second) before Grover overtakes brute-force searching. Shor’s quantum algorithms for factoring and discrete logarithm \[Sho97\] give us better examples. Here we achieve an exponential speed-up from the best known probabilistic algorithms. Factoring alone does not make quantum computing useful. Consider the following paragraph from Shor’s famous paper \[Sho97, p. 1506\]: > Discrete logarithms and factoring are not in themselves widely useful problems. They have become useful only because they have been found to be crucial for public-key cryptography, and this application is in turn possible only because they have been presumed to be difficult. This is also true of the generalizations of Boneh and Lipton \[BL95\] of these algorithms. If the only uses of quantum computation remain discrete logarithms and factoring, it will likely become a special-purpose technique whose only raison d’etre is to thwart public key cryptosystems. How useful can quantum computing be? What is the computational complexity of $`\mathrm{𝐁𝐐𝐏}`$, the class of efficiently quantum computable problems. Does $`\mathrm{𝐁𝐐𝐏}`$ contain $`\mathrm{𝐍𝐏}`$-complete problems like graph 3-colorability? What is the relationship of $`\mathrm{𝐁𝐐𝐏}`$ and other more traditional complexity classes like the polynomial-time hierarchy? These questions are well suited for computational complexity theorists. Many researchers though shy away from quantum complexity as it appears that a deep knowledge of physics is necessary to understand $`\mathrm{𝐁𝐐𝐏}`$. Also most papers use a strange form of notation that make their results appear more difficult than they really are. This paper argues that one can think about $`\mathrm{𝐁𝐐𝐏}`$ as just a regular complexity class, not conceptually much different than its probabilistic cousin $`\mathrm{𝐁𝐏𝐏}`$. We develop a way of looking at computation as matrix multiplications and exhibit some immediate effects of this formulation. We show that $`\mathrm{𝐁𝐐𝐏}`$ is a rather robust complexity class that has much in common with other classes. We argue that $`\mathrm{𝐁𝐐𝐏}`$ and $`\mathrm{𝐁𝐏𝐏}`$ exhibit much of the same behavior particularly in their ability to do searching and hiding of information. While this paper does not assume any knowledge of physics, we do recommend a familiarity with basic notions of complexity theory as described in Balcázar, Díaz and Gabarró \[BDG88\]. ## 2 Computation as Matrix Multiplication Consider a multitape deterministic Turing machine $`M`$ that uses $`t(n)`$ time and $`s(n)`$ space. Let us also fix some input $`x`$ in $`\{0,1\}^n`$ and let $`t=t(n)`$ and $`s=s(n)`$. We will assume $`t(n)n`$ and $`s(n)\mathrm{log}n`$ are fully time and space constructible respectively. We will always have $`t(n)s(n)2^{O(t(n))}`$. Recall that a configuration $`c`$ of machine $`M`$ on input $`x`$ consists of the contents of the work tapes, tape pointers and current state. We will let $`C`$ be the number of all configurations of $`M`$ on input $`x`$. Note that a read-only input tape is not considered as part of the configuration and thus we have $`C=2^{O(s)}`$. For most of the paper, we will refer to the polynomial-time case, where $`t(n)=n^{O(1)}`$ and $`s(n)=2^{t(n)}=2^{n^{O(1)}}`$. Fix an input $`x`$. Let $`c_I`$ be the initial configuration of $`M(x)`$. We can assume without loss of generality that $`M`$ has exactly one accepting configuration $`c_A`$ and that once $`M`$ enters $`c_A`$ it remains in $`c_A`$. Define a $`C\times C`$ transition matrix $`T`$ from the description of the transition function of $`M`$. Let $`T(c_a,c_b)=1`$ if one can get from $`c_a`$ to $`c_b`$ in one step and $`T(c_a,c_b)=0`$ otherwise. ###### Observation 2.1 For any two configurations $`c_a`$ and $`c_b`$, $`T^r(c_a,c_b)=1`$ if and only if $`M`$ on input $`x`$ in configuration $`c_a`$ when run for $`r`$ steps will be in configuration $`c_b`$. In particular we have that $`M(x)`$ accepts if and only if $`T^t(c_I,c_A)=1`$. Now what happens if we allow $`M`$ to be a nondeterministic machine. Let us formally describe how the transition function $`\delta `$ works for a $`k`$-tape machine (along the lines of Hopcroft and Ullman \[HU79\]). $$\delta :Q\times \mathrm{\Sigma }\times \mathrm{\Gamma }^k\times Q\times \mathrm{\Gamma }^k\times \{L,R\}^{k+1}\{0,1\}$$ Here $`Q`$ is the set of states, $`\mathrm{\Sigma }`$ is the input alphabet and $`\mathrm{\Gamma }`$ the tape alphabet. The transition function $`\delta (q,a,b_1,\mathrm{},b_k,p,c_1,\mathrm{},c_k,d_0,d_1,\mathrm{},d_k)`$ takes the value one if the transition from state $`q`$ looking at input bit $`a`$ and work tape bits $`b_1`$, $`b_2`$, $`\mathrm{}`$, $`b_k`$ moving to state $`p`$, writing $`c_1`$, $`\mathrm{}`$, $`c_k`$ on the work tapes, moving the input head in the direction $`d_0`$ and the tape heads in directions $`d_1`$, $`\mathrm{}`$, $`d_k`$ respectively is legal. This $`\delta `$ function yields again a transition matrix $`T`$ in the obvious way: If it is possible to reach $`c_b`$ from $`c_a`$ in one step then $`T(c_a,c_b)=1`$. If it is not possible to go from $`c_a`$ to $`c_b`$ in one transition then $`T(c_a,c_b)=0`$. A few notes about the matrix $`T`$: Most entries are zero. For polynomial-time computation, given $`c_a`$, $`c_b`$ and $`x`$ one can in polynomial-time compute $`T(c_a,c_b)`$, even though the whole transition matrix is too big to write down in polynomial-time. Observation 2.1 has a simple generalization: ###### Observation 2.2 For any two configurations $`c_a`$ and $`c_b`$, $`T^r(c_a,c_b)`$ is the number of computation paths from $`c_a`$ to $`c_b`$ of length $`r`$. We now have that $`M`$ on input $`x`$ accepts if and only if $`T^t(c_I,c_A)>0`$. We can define $`\mathrm{\#}M(x)=T^t(c_I,c_A)`$ as the number of accepting paths of machine $`M`$ on input $`x`$. For polynomial-time machines this yields the $`\mathrm{\#}𝐏`$ functions first defined by Valiant \[Val79\]. Now interesting events happen when we allow our $`\delta `$ function to take on nonbinary values. First let us consider $`\delta `$ taking on nonnegative integers. We get a further generalization of Observations 2.1 and 2.2: ###### Observation 2.3 The value of $`T^r(c_a,c_b)`$ is the sum over all computation paths from $`c_a`$ to $`c_b`$ of the product of the values of $`\delta `$ over each transition on the path. What kind of functions do we get from $`T^t(c_I,c_A)`$ for polynomial-time machines? We still get exactly the $`\mathrm{\#}𝐏`$ functions. Suppose that we have a $`T(c_a,c_b)=k`$. We simply create a new nondeterministic Turing machine that will have $`k`$ computation paths, of length about $`\mathrm{log}k`$, from $`c_a`$ to $`c_b`$. This will only increase the running time by a constant factor. Now suppose we allow our $`\delta `$ function to take on nonnegative rational values. Remember that the $`\delta `$ function is a finite object independent of the input. Let $`v`$ be the least common multiple of the denominators of all the possible values of the $`\delta `$ function. Define $`\delta _{}=v\delta `$. Let $`T`$ and $`T_{}`$ be the corresponding matrices for $`\delta `$ and $`\delta _{}`$ respective. Note $`T_{}=vT`$. The values $`T_{}^t`$ still capture the $`\mathrm{\#}𝐏`$ functions. We also have $`T^t=T_{}^t/v^t`$. Since $`v^t`$ is easily computable this does not give us much more power than before. We can use a restricted version of this model to capture probabilistic machines. Probabilistic machines have $$\delta :Q\times \mathrm{\Sigma }\times \mathrm{\Gamma }^k\times Q\times \mathrm{\Gamma }^k\times \{L,R\}^{k+1}[0,1]$$ with the added restriction that for any fixed values of $`q`$, $`a`$ and $`b_1,\mathrm{},b_k`$, $$\underset{p,c_1,\mathrm{},c_k,d_0,d_1,\mathrm{},d_k}{}\delta (q,a,b_1,\mathrm{},b_k,p,c_1,\mathrm{},c_k,d_0,d_1,\mathrm{},d_k)=1.$$ What does this do to the matrices $`T`$? Every row of $`T`$ sums up to exactly one and its entries are all range in between zero and one, i.e., the matrices are stochastic. Note also that stochastic matrices preserve the $`L_1`$ norm, i.e., for every vector $`u`$, $`L_1(uT)=L_1(u)`$. These observations will be useful to us as we try to understand quantum computation. Now $`T^t(c_I,c_A)`$ computes exactly the probability of acceptance of our probabilistic machine. Keep in mind that $`T`$ still is mostly zero and although the matrix $`T`$ is exponential size, the entries are easily computable given the indices. We can define the class $`\mathrm{𝐁𝐏𝐏}`$: A language $`L`$ is in $`\mathrm{𝐁𝐏𝐏}`$ if there is a probabilistic matrix $`T`$ as described above and a polynomial $`t`$ such that * For $`x`$ in $`L`$, we have $`T^t(c_I,c_A)2/3`$. * For $`x`$ not in $`L`$, we have $`T^t(c_I,c_A)1/3`$. Using standard repetition tricks to reduce the error we can replace $`2/3`$ and $`1/3`$ above with $`12^{p(|x|)}`$ and $`2^{p(|x|)}`$ for any polynomial $`p`$. Let us make some variations to the probabilistic model. First let us allow $`\delta `$ to take on arbitrary rational values including negative values. Since the value $`T^t`$ may be negative we will use $`(T^t(c_I,c_A))^2`$ as our “probability of acceptance”. Finally instead of preserving the $`L_1`$ norm, we preserve the $`L_2`$ norm instead. That is for every vector $`u`$, we have $$L_2(T(u))=L_2(u)=\sqrt{\underset{a}{}u_a^2}.$$ Unlike the $`L_1`$ norm, preserving the $`L_2`$ norm is a global condition, we cannot just require that the squares of the values of the $`\delta `$ function add up to one. Matrices that preserve the $`L_2`$ norm are called unitary. This model looks strange but still rather simple. Yet this model captures exactly the power of quantum computing! We can define the class $`\mathrm{𝐁𝐐𝐏}`$: A language $`L`$ is in $`\mathrm{𝐁𝐐𝐏}`$ if there is a quantum matrix $`T`$ as described above and a polynomial $`t`$ such that * For $`x`$ in $`L`$, we have $`(T^t(c_I,c_A))^22/3`$. * For $`x`$ not in $`L`$, we have $`(T^t(c_I,c_A))^21/3`$. ## 3 Questions For those who have seen quantum computing talks in the past, many questions may come up. ### 3.1 Where’s the physics? The presentation of quantum computing above is based mostly on the computation model developed by Bernstein and Vazirani \[BV97\]. We have presented the model here as another complexity model. It does however capture all of the physical power and rules of quantum computation. Can one study quantum computation without a deep background in quantum mechanics? I say yes. I drive a car without much of a clue as to what a carburetor does. I can program a classical computer and do research on the computational complexity of these computers even though I do not have a real understanding of how a transistor works. I often do research on theoretical computation models such as nondeterminism that have no physical counterpart at all. Given a good model of a quantum computer, computer scientists can study its computational abilities without much knowledge of the physical properties of that model. ### 3.2 Where’s the $``$bra$`|`$ and $`|`$ket$``$ notation? When researchers in two disciplines try to work they usually find that language forms a major wall between them. Not the base spoken language, but the notation and definitions each one assumes that every first-year graduate student in the field knows and follows. In quantum computing, generally physics notation has prevailed, leading to one of the major barriers to entering the area. Physicists generally use the Dirac bra-ket notation to represent base quantum states. When computer scientists see this notation, it looks quite foreign and many assume that the complexity of the notation reflects some deep principles in quantum mechanics. Nothing could be further from the truth as the bra-ket notation represents nothing more than vectors. The notation $`|010`$ is called a “ket” and represents the string $`010`$. In general $`|x`$ with $`x\{0,1\}^k`$ represents the string $`x`$. Concatenation is concatenation: $`|x|y=|xy`$. The strings of length $`k`$ form the basis of a vector space of dimension $`2^k`$. These basis vectors roughly correspond to the configurations described in Section 2. Now, there is also a “bra” and various rules that apply when putting bras and kets together but these rarely come into play in quantum computing. For the interested reader there are many sources one can read on the subject such as Preskill’s lecture notes \[Pre98\]. Why does the simple looking notation cause such confusion for computer scientists? This notation does not conform to computer science conventions, particular to that of symmetry. When we consider vectors or other groupings of objects, computer scientists always use symmetric brackets such at $`8,3,4`$, $`[0\mathrm{}1]`$, $`(a+b)^{}`$ and $`|x|`$. When typical computer scientists look at a character like “$``$” or “$`|`$” that does not have a counterpart, they consider these as relational operators, making equations like $`|\psi _1(y)={\displaystyle \frac{1}{\sqrt{2^n}}}{\displaystyle \underset{x}{}}(1)^{\delta _{x,y}}|x=|\psi _0{\displaystyle \frac{2}{\sqrt{2^n}}}|y`$ \[BBBV97, p. 1514\] (1) impossible even to parse. As this paper suggests, one does not need to know Dirac notation to understand the basics of quantum computation. Unfortunately one is forced to learn it to follow the communication in the field. If theoretical computer scientists collaborate then they must learn . It does not matter if they like some other mathematical document software, they have to use if they wish to write papers with other computer scientists. Likewise if you want to truly study quantum computation you will need to learn the Dirac notation. One can easily develop more natural notation for computer scientists but we have already lost the battle. Eventually the notation becomes easier to parse and follow and formulas like Equation (1) seem almost normal. I find it a shame though that computer scientists find it difficult to go to talks on quantum computing not so much because of the complexity of the subject matter but because of the abnormal notation. ### 3.3 Don’t we need arbitrary real and complex numbers to do quantum computation? In short no. The physics definition allows arbitrary real and complex entries as long as the matrices are unitary. This by itself limits the values to have absolute value at most one. But one can do much better. If we allow all possible reals, $`\mathrm{𝐁𝐐𝐏}`$ can accept arbitrarily complicated languages \[ADH97\]. However this result feels like cheating—basically you encode hard languages directly into the entries of the matrix. Thus one requires knowing the language ahead of time to create the machine. A similar trick can also be played with probabilistic machines using noncomputable probabilities. For fairness we should only allow efficiently computable matrix entries, where we can compute the $`i`$th bit in time polynomial in $`i`$. Independently Adleman, DeMarrais and Huang \[ADH97\] and Solovay and Yao \[SY96\] show that we can simulate a $`\mathrm{𝐁𝐐𝐏}`$ machine using efficiently computable entries from the set $`\{1,\frac{4}{5},\frac{3}{5},0,\frac{3}{5},\frac{4}{5},1\}`$. Or you can get away with fewer numbers if you don’t mind an irrational: $`\{1,\sqrt{2},0,\sqrt{2},1\}`$. ### 3.4 Don’t we have to require the computation to be reversible? The matrix $`T`$ as well as any power of $`T`$ is unitary, i.e. it preserves the $`L_2`$ norm. For real valued matrices, $`T`$ is unitary if and only if the transpose of $`T`$ is the inverse of $`T`$. The reader should try the proof himself or it can be found in Bernstein and Vazirani \[BV97, p. 1463\]. In particular this means that $`T`$ is invertible. Even more so the inverse of $`T`$ is itself unitary. If you have a vector $`v`$ over the configurations that gives the value for each configuration, let $`w=T(v)`$. We then have $`T^1(w)=v`$. Given the current entire state of the system we can reverse it to get the previous state. Don’t think of reversibility as a requirement of quantum computing rather as a consequence. Nevertheless keep in mind that in creating a quantum algorithm at a minimum you will need to create a reversible procedure. ### 3.5 What if there are many accepting configurations? For traditional models of computation we can usually assume only one accepting configuration. We do this by a cleanup operation—after the machine decides to accept we erase the work tapes, move the head to the left and enter a specified final accepting state. This procedure will not work for quantum computation since it is not reversible. Bennett, Bernstein, Brassard and Vazirani \[BBBV97\] show an interesting way to get around this problem: Compute whether the quantum machine accepts. Copy the answer to an extra quantum bit. Then reverse the whole process. The unique accepting state is now the initial state with this extra bit turned on. ### 3.6 Don’t we have to allow measurements? Our formulation of $`\mathrm{𝐁𝐐𝐏}`$ does not appear to contain any measurements. In fact we are simulating a measurement at the end. A $`\mathrm{𝐁𝐐𝐏}`$ machine will output “accept” with probability equal to the sum of the squares of the values of the accepting configurations. We have only one accepting configuration so we need just consider $`(T^t(c_I,c_A))^2`$. One could consider a process that allows measurements during the computation. Bernstein and Vazirani \[BV97\] show that we can push all of the measurements to the end. Basically measurements are linear projections and we can simulate a measurement by doing a rotation and taking one big projection at the end. ## 4 Full Robustness Section 3 points to the fact that $`\mathrm{𝐁𝐐𝐏}`$ is quite a robust complexity class. Bernstein and Vazirani \[BV97\] show two more important facts: That $`\mathrm{𝐁𝐐𝐏}`$ can simulate any deterministic or probabilistic polynomial-time algorithm and $`\mathrm{𝐁𝐐𝐏}^{\mathrm{𝐁𝐐𝐏}}=\mathrm{𝐁𝐐𝐏}`$. In other words we can do quantum subroutines and build that directly into the quantum computation. The proofs of these facts are technically quite complicated but here a few points to think about. Note that a deterministic computation might not be reversible but Bennett \[Ben89\] showed how to simulate any deterministic computation by a reversible computation with only squaring the computation time. Flipping a truly random coin seems inherently nonreversible: Once flipped how does one recover the previous state? Quantum computing does have one nonreversible operation known as measurements. One can simulate truly random coin tosses using measurements, or as described in Section 3.6 simulating a measurement and doing it at the end. From these results we can use that standard error reduction techniques to reduce the error in a quantum algorithm to $`2^{p(n)}`$ where $`p`$ is a polynomial and $`n`$ is in the length of the input. Other complexity classes such as the zero-error version of $`\mathrm{𝐁𝐐𝐏}`$ (denoted $`\mathrm{𝐄𝐐𝐏}`$ or $`\mathrm{𝐐𝐏}`$) may not have some of the nice robust properties of Section 3 since approximating the matrix entries would no longer keep us in the class. Space-bounded classes also lack some of the robustness of $`\mathrm{𝐁𝐐𝐏}`$. For example if we do not allow a measurement until the end even simulating coin flips becomes difficult: we cannot reuse the coins and keep reversibility. See Watrous \[Wat99b\] for a discussion. ## 5 Using This Formulation Once you have the formulation of quantum machines described in Section 2, one immediately gets interesting results on quantum complexity. ### 5.1 $`\mathrm{𝐀𝐖𝐏𝐏}`$ Suppose we remove the unitary restriction in the definition of $`\mathrm{𝐁𝐐𝐏}`$ in Section 2. This yields the class $`\mathrm{𝐀𝐖𝐏𝐏}`$. Li \[Li93\] and Fenner, Fortnow, Kurtz and Li \[FFKL93\] defined and studied the class $`\mathrm{𝐀𝐖𝐏𝐏}`$ (stands for Almost-Wide Probabilistic Polynomial-time) extensively. They showed many interesting results for this class. 1. $`\mathrm{𝐀𝐖𝐏𝐏}\mathrm{𝐏𝐏}\mathrm{𝐏𝐒𝐏𝐀𝐂𝐄}`$, where $`\mathrm{𝐏𝐏}`$ is the set of languages accepted by a probabilistic polynomial-time Turing machine where we only require the error probability to be smaller than one-half. 2. The class $`\mathrm{𝐀𝐖𝐏𝐏}`$ is low for $`\mathrm{𝐏𝐏}`$, i.e., for any language $`L`$ in $`\mathrm{𝐀𝐖𝐏𝐏}`$, $`\mathrm{𝐏𝐏}^L=\mathrm{𝐏𝐏}`$. 3. If $`𝐏=\mathrm{𝐏𝐒𝐏𝐀𝐂𝐄}`$ then relative to any generic oracle $`G`$, $`𝐏^G=\mathrm{𝐀𝐖𝐏𝐏}^G`$. This implies that without any assumption there is a relativized world $`A`$ such that $`𝐏^A=\mathrm{𝐀𝐖𝐏𝐏}^A`$ and the polynomial-time hierarchy is infinite. 4. There exists a relativized world where $`\mathrm{𝐀𝐖𝐏𝐏}`$ does not have complete sets, in fact where $`\mathrm{𝐀𝐖𝐏𝐏}`$ has no sets hard for all of $`\mathrm{𝐁𝐏𝐏}`$. Relativization results show us important limitations on how certain techniques will solve problems in complexity theory. For a background on relativization see the survey paper by Fortnow \[For94\]. Fortnow and Rogers \[FR99\] observed that $`\mathrm{𝐁𝐐𝐏}\mathrm{𝐀𝐖𝐏𝐏}`$ basically falls out of the characterization given in Section 2. From this we have several corollaries. 1. $`\mathrm{𝐁𝐐𝐏}\mathrm{𝐏𝐏}\mathrm{𝐏𝐒𝐏𝐀𝐂𝐄}`$ \[ADH97\] 2. The class $`\mathrm{𝐁𝐐𝐏}`$ is low for $`\mathrm{𝐏𝐏}`$. 3. There exists a relativized world where $`𝐏=\mathrm{𝐁𝐐𝐏}`$ and the polynomial-time hierarchy is infinite. 4. There exists a relativized world where $`\mathrm{𝐁𝐐𝐏}`$ has no complete sets, not even any sets hard for $`\mathrm{𝐁𝐏𝐏}`$. ### 5.2 Decision Trees The oracle results mentioned in Section 5.1 follow from results on decision tree complexity that have interest in their own right. Consider the situation where we wish to compute a function from $`\{0,1\}^N`$ to $`\{0,1\}`$ where access to the input is limited as oracle questions. Here the complexity measure is the number of queries needed to compute the function, not the running time. Beals, Buhrman, Cleve, Mosca and de Wolf \[BBC<sup>+</sup>98\] show how to vary the matrix model of Section 2 to create special linear transformations that describe a reversible oracle query. One can interleaves matrices describing these transformations with the unitary matrices given by the computation of the Turing machine. Grover \[Gro96\] shows how to get a nontrivial advantage with quantum computers: He shows that computing the OR function needs only $`O(\sqrt{N})`$ queries although classically $`\mathrm{\Omega }(N)`$ input bits are needed in the worst case. Bernstein and Vazirani \[BV97\] give a superpolynomial gap and Simon \[Sim97\] gives an exponential gap. Both of these gaps require that there are particular subsets of the inputs to which $`f`$ is restricted. Beals, Buhrman, Cleve, Mosca and de Wolf \[BBC<sup>+</sup>98\] notice that the characterization given in Section 2 shows that any quantum decision tree algorithm making $`t`$ queries is a polynomial in the queries of degree at most $`2t`$. Nisan and Szegedy \[NS94\] show that if a polynomial of degree $`d`$ on $`\{0,1\}^N`$ approximates the characteristic function of some language $`L\{0,1\}^N`$ then $`L`$ has deterministic decision tree complexity polynomial in $`d`$. Combining these ideas, if there is a quantum algorithm computing a function $`f`$ defined on all of $`\{0,1\}^N`$ and using $`t`$ queries then there exists a deterministic algorithm computing $`f`$ using polynomial in $`t`$ queries. ## 6 Probabilistic versus Quantum Often one hears of the various ways that quantum machines have advantages over classical computation, for example, the ability to go into many parallel states and that states can be entangled with each other. In fact often these properties occur in probabilistic computation as well. The strength of quantum computing lies in the ability to have bad computation paths eliminate each other thus causing some good paths to occur with larger probability. The ability for quantum machines to take advantage of this interference is tempered by their restriction to unitary operations. Suppose you videotape a football game from the television. Let us say the teams are evenly matched so that either team has an equal probability of winning. Now make a copy of the tape without watching it. Now these videotapes are in some sense entangled. We don’t know who won the game until we watch it but both outcomes will be the same. I can take a videotape to Mars and watch it there and at the end I will instantaneously know the outcome of the game on the other tape. Entanglement works quite differently for quantum computing. In particular the videotape model fails to capture quantum entanglement—no “hidden variable” theory can capture the effects of quantum entanglement (see a discussion in Preskill \[Pre98\]). However we should not consider entanglement to be of much help in the power of $`\mathrm{𝐁𝐐𝐏}`$. When it comes to parallelism there is really no difference between probabilistic and quantum computation. Consider the value $`T^t(c_I,c_A)`$ expanded as follows: $$T^t(c_I,c_A)=\underset{c_1,\mathrm{},c_{t1}}{}T(c_I,c_1)T(c_1,c_2)\mathrm{}T(c_{t2},c_{t1})T(c_{t1},c_A)$$ One can think of the vector $`c_1,\mathrm{},c_{t1}`$ as describing a computation path, the value $$T(c_I,c_1)T(c_1,c_2)\mathrm{}T(c_{t2},c_{t1})T(c_{t1},c_A)$$ as the value along that path and the final value as the sum of the values over all paths. Did this view of parallelism depend on whether we considered quantum or probabilistic computation? Not a bit. Where does the power of quantum come from? From interference. For probabilistic computation the value $$T(c_I,c_1)T(c_1,c_2)\mathrm{}T(c_{t2},c_{t1})T(c_{t1},c_A)$$ is always nonnegative. Once a computation path has a positive value it is there to stay. For quantum computation the value could be negative on some path causing cancellation or interference. This allows other paths to occur with a higher probability. The restriction of quantum computers to unitary transformations limits the applications of interference but still we can achieve some useful power from it. ### 6.1 Searching versus Hiding Randomness plays two important but quite different roles in the theory of computation. Typically we use randomness for searching. In this scenario, we have a large search space full of mostly good solutions. Deterministically we can only look in a relatively small number of places and may miss the vast majority of good solutions. By using randomness we can, with high probability, find the good solutions. Typical examples include primality testing \[SS77\] and searching undirected graphs in a small amount of space \[AKL<sup>+</sup>79\]. However these randomness techniques appear to apply to only a small number of problems. Recent results in derandomization indicate we probably can do as well with deterministically chosen pseudorandom sequences. We also use randomness to hide information. A computer, no matter how powerful, cannot predict the value of a truly random bit it cannot see. Hiding forms the basis of virtually all cryptographic protocols. Hiding also plays an important role in complexity theory. The surprising strength of interactive proof systems \[LFKN92, Sha92, BFL91, ALM<sup>+</sup>92\] comes from the inability of the prover to predict the verifier’s future coin tosses. We have the same dichotomy between searching and hiding in quantum bits. For quantum searching we do not necessarily require that most paths have high probability. The use of quantum interference allows us to find some things with certain kinds of structure. Shor’s algorithms for factoring and discrete logarithm \[Sho97\] form the obvious example here. While we do not expect that there exists any notion of quantum pseudorandom generators, at the moment we still only have a limited collection of problems where quantum search greatly helps over classical methods. Quantum hiding, like its probabilistic counterpart, has a powerful effect on computation theory. Not only can no computer predict the value of a quantum bit not yet measured but the quantum bit cannot be copied and any attempt at early measurement will change the state of the quantum bit. Quantum bits do have one disadvantage over probabilistic bits. When we flip a probabilistic coin the original state has been irrevocably destroyed. The reversibility of quantum computation prevents destroying quantum state. However, we can get around this problem by using classical probabilistic bits to determine how to prepare the quantum states. We can always get classical probabilistic bits by measuring an appropriate quantum state. Quantum hiding in this manner gives powerful tools for quantum cryptography \[BBE92\] and quantum interactive proof systems \[Wat99a\] that go beyond what we believe possible with classical randomness. ## 7 Future Directions Quantum computation is ripe for complexity theorists. There are still many fundamental questions remaining to be solved. Is $`\mathrm{𝐍𝐏}`$ in $`\mathrm{𝐁𝐐𝐏}`$? While we probably cannot answer that problem directly we could show some unlikely consequences like the polynomial-time hierarchy collapses. Does $`\mathrm{𝐁𝐐𝐏}`$ sit in the polynomial-time hierarchy? The only evidence against this comes from Bernstein-Vazirani \[BV97\]. In their paper they create a relativized world $`A`$ and a language $`L`$ that sits in $`\mathrm{𝐄𝐐𝐏}^A\mathrm{𝐁𝐐𝐏}^A`$ but not $`\mathrm{𝐁𝐏𝐏}^A`$. We cannot prove that $`L`$ sits in the polynomial-hierarchy relative to $`A`$. However the language $`L`$ does sit in deterministic quasipolynomial ($`2^{\mathrm{log}^{O(1)}n}`$) time and by an alternating Turing machine that uses only polylogarithmic alternations where these machines have access to $`A`$. So while we may find it difficult to show $`\mathrm{𝐁𝐐𝐏}`$ is in the polynomial-time hierarchy, we could show that every language in $`\mathrm{𝐁𝐐𝐏}`$ is accepted by some alternating Turing machine using polylogarithmic alternations and/or polynomial-time. For such results we will need more than the fact than $`\mathrm{𝐁𝐐𝐏}`$ sits in $`\mathrm{𝐀𝐖𝐏𝐏}`$ as described in Section 5.1. Relative to random oracles $`\mathrm{𝐍𝐏}`$ is in $`\mathrm{𝐀𝐖𝐏𝐏}`$ and there exist relative worlds where an infinite polynomial-time hierarchy sits strictly inside of $`\mathrm{𝐀𝐖𝐏𝐏}`$. Somehow we will need to make better use of the unitary nature of the quantum transformations. Shor’s algorithm shows that quantum machines can defeat many commonly used cryptographic protocols. Can we create one-way functions that no quantum machine can invert better than random guessing? Perhaps we will need such functions computed by quantum machines as well. Maybe quantum relates to other classes out there? Is $`\mathrm{𝐁𝐐𝐏}`$ equivalent to the problems having statistical zero-knowledge proof systems? This seems far fetched but we have no evidence against this. For further reading I recommend that the reader get a hold of SIAM Journal on Computing Volume 26 Number 5 (1997). Half of this issue is devoted to quantum computation and as one can see from the references it has many of the important papers in the area. Preskill’s course notes \[Pre98\] give a detailed description of quantum computation from the physicists’ point of view. Books on the quantum computation are also starting to appear. An early book by Gruska \[Gru99\] gives a broad range of work on quantum computation. ## Acknowledgments I wish to thank the organizers of the AQIP ’99 conference for inviting me to give the presentation and inviting me to write this paper based on that presentation for the special issue of the conference. I have had many exciting discussions on these topics with several top researchers in the field including Harry Buhrman, Peter Shor, Charlie Bennett, Gilles Brassard and Dorit Aharanov. The opinions expressed in this paper remain solely my own. Dieter van Melkebeek gave many useful comments on an earlier draft of this paper.
warning/0003/astro-ph0003119.html
ar5iv
text
# 1 Introduction ## 1 Introduction The nearby FR-I radio galaxy NGC 4261 (3C270) is a good candidate for the detection of free-free absorption by ionized gas in an inner accretion disk. The galaxy is known to contain a central black hole with a mass of $`5\times 10^8\mathrm{M}_{}`$, a nearly edge-on nuclear disk of gas and dust with a diameter of $`100`$ pc, and a large-scale symmetric radio structure which implies that the radio axis is close to the plane of the sky. At an assumed distance of 40 Mpc, 1 milliarcsecond (mas) corresponds to 0.2 pc. Previous VLBA observations of this galaxy revealed a parsec-scale radio jet and counterjet aligned with the kpc-scale jet (see Figure 1). The opening angle of the jets is less than $`20^{}`$ during the first 0.2 pc and $`<5^{}`$ during the first 0.8 pc. At 8.4 GHz we found evidence for a narrow gap in radio brightness at the base of the parsec-scale counterjet, just east (left) of the brightest peak which we identified as the core based on its inverted spectrum between 1.6 and 8.4 GHz (see the left part of Figure 2, from Jones and Wehrle 1997). We tentatively identified this gap as the signature of free-free absorption by a nearly-edge on inner disk with a width $`<<0.1`$ pc and an average electron density of $`10^310^8\mathrm{cm}^3`$ over the inner 0.1 pc. ## 2 Observations We observed NGC 4261 at 1.6 and 4.9 GHz with HALCA and a ground array composed of 7 VLBA antennas plus Shanghai, Kashima, and the DSN 70-m Tidbinbilla antennas at 1.6 GHz (22 June 1999) and 8 VLBA antennas plus the phased VLA at 4.9 GHz (27 June 1999). During both epochs the VLBA antennas at St. Croix and Hancock were unavailable, as was the North Liberty antenna at 1.6 GHz. Data were recorded as two 16-MHz bandwidth channels with 2-bit sampling by the Mark-III/VLBA systems and correlated at the VLBA processor in Socorro. Both channels were sensitive to left circular polarization. Fringe-fitting was carried out in AIPS after applying a priori amplitude calibration. For VLBA antennas we used continuously measured system temperatures, for the VLA we used measured $`\mathrm{T}_\mathrm{A}/\mathrm{T}_{\mathrm{SYS}}`$ values with an assumed source flux density of 5 Jy, and for the remaining antennas we used typical gain and system temperature values obtained from the VSOP web site. Fringes were found to all antennas at 1.6 GHz except HALCA, but the signal/noise ratio to Shanghai and Kashima was very low and these data were not used. The a priori amplitude calibration for Tidbinbilla was dramatically incorrect for unknown reasons. We calibrated Tidbinbilla by imaging the compact structure of the source using VLBA data, then holding the VLBA antenna gains fixed and allowing the Tidbinbilla gain to vary. This produced a good match in correlated flux density where the projected VLBA and Tidbinbilla baselines overlap. At 4.9 GHz fringes were found to all antennas, including HALCA. A similar correction to the a priori amplitude calibration for HALCA and the phased VLA was applied. In both observations we found that averaging in frequency over both 16-MHz channels in AIPS produced large, baseline-dependent amplitude reductions even though the post-fringe-fit visibility phases were flat and continuous between channels. Averaging over frequency within each 16-MHz band separately fixed this problem. Difmap was used for detailed data editing, self-calibration, and image deconvolution. Both 16-MHz bands were combined during imaging. Imaging within Difmap used uniform weighting with the weight of HALCA data increased by a factor of 500. Several iterations of phase-only self calibration, followed by amplitude self calibration iterations with decreasing time scales, resulted in good fits ($`\chi ^21`$) between the source model and the data. ## 3 Results ### 3.1 1.6 GHz Although our image at 1.6 GHz does not include data from HALCA, it does have more than twice the angular resolution of our previous 1.6 GHz image (Jones and Wehrle 1997) due to the addition of Tidbinbilla. The previous image showed a symmetric structure, with the jet and counterjet extending west and east from the core. No evidence for absorption is seen in this image. However, with higher resolution we do detect a narrow gap in emission just east of the core, at the base of the counterjet. The width of the gap is less than 2 mas. ### 3.2 4.9 GHz We detected fringes to HALCA at 4.9 GHz when the projected Earth-space baselines were less than one Earth diameter. The HALCA data fills in the (u,v) coverage hole between continental VLBA baselines and those to Mauna Kea, and also increases the north-south resolution by a factor of two. Our 4.9 GHz image is shown in Figure 2. Note that the gap in emission is again seen just east of the peak. A careful comparison of brightness along the radio axis at 4.9 and 8.4 GHz shows that the gap is both deeper and wider at 4.9 GHz, as expected from free-free absorption. The region of the gap has a very inverted spectrum, the brightest peak (core) has a slightly less inverted spectrum, and the distant parts of both the jet and counterjet have steep spectra. ## 4 Summary Our observations at 1.6 and 4.9 GHz appear to confirm the free-free absorption explanation for the sub-parsec radio morphology in NGC 4261. Measurements of the optical depth in the absorbed region and the distance between the absorption and the core as a function of frequency will allow the radial distribution of electron density in the inner parsec of the disk to be determined. #### Acknowledgements. We gratefully acknowledge the VSOP Pro-ject, which is led by the Japanese Institute of Space and Astronautical Science in cooperation with many organizations and radio telescopes around the world. This research was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the U.S. National Aeronautics and Space Administration. ## References Jones, D.L. & Wehrle, A.E. 1997, ApJ, 484, 186
warning/0003/hep-ph0003212.html
ar5iv
text
# Neutrino Masses and Mixing at the Millennium*footnote **footnote *Talk presented at the 5th International Conference on Physics Potential & Development of μ⁺⁢μ⁻ Colliders, San Francisco, December 1999. ## I Neutrino Counting How many neutrinos are there? Neutrino counting at LEP of $`Z\nu \overline{\nu }`$ decays obtains $`N_\nu =3`$ active flavors — the expected $`\nu _e,\nu _\mu ,\nu _\tau `$. However, light isosinglet, right-handed “sterile” neutrinos with no gauge boson interactions could also exist. Big Bang Nucleosynthesis determines the equivalent number of massless neutrinos at the time of nucleosynthesis. The bounds inferred ($`N_\nu <3.2`$burles , $`N_\nu <4`$lisi , $`N_\nu <5.3`$olive ) depend on which measurements of the <sup>4</sup>He and D/H abundances are used in the analysis. Neutrino oscillation phenomenology for $`N_\nu =3`$ and $`N_\nu =4`$ is very different and both options need to be considered at present. ## II Neutrino Masses Tree-level mass generation occurs through the Higgs mechanism. The Dirac mass $`m_D`$ arises in a lepton conserving ($`\mathrm{\Delta }L=0`$) interaction (see Fig. 1) and requires a right-handed neutrino. A Majorana mass $`m_M`$ occurs through a $`\mathrm{\Delta }L=2`$ process with only a left-handed light neutrino field and a heavy isosinglet intermediate field $`N^c`$. Then the see-saw mechanismseesaw with $`m_D10^2`$ GeV and $`m_M>10^{12}`$ GeV generates light neutrinos $$m_\nu =\frac{m_D^2}{m_M}$$ (1) that are nearly Majorana. In the case that $`m_Dm_M\mathrm{eV}`$, as can be realized in some modelslangacker ; BLLP , active–sterile neutrino oscillations can take place. Neutrino mass can alternatively be generated radiatively by new interactionsRMreview , such as the $`R`$-parity violating $`\nu b\stackrel{~}{b}`$ interactiondrees , as illustrated in Fig. 2. ## III Three-Neutrino Oscillatons The relation of the three-neutrino flavor eigenstates to the mass eigenstates is $$\left(\begin{array}{c}\nu _e\\ \nu _\mu \\ \mu _\tau \end{array}\right)=U\left(\begin{array}{c}\nu _1\\ \nu _2\\ \nu _3\end{array}\right),$$ (2) where $`U`$ is the $`3\times 3`$ Maki-Nakagawa-Sakata (MNS) mixing matrixMNS . It can be parametrized by $`U`$ $`=`$ $`\left(\begin{array}{ccc}c_{13}c_{12}& c_{13}s_{12}& s_{13}e^{i\delta }\\ c_{23}s_{12}s_{13}s_{23}c_{12}e^{i\delta }& c_{23}c_{12}s_{13}s_{23}s_{12}e^{i\delta }& c_{13}s_{23}\\ s_{23}s_{12}s_{13}c_{23}c_{12}e^{i\delta }& s_{23}c_{12}s_{13}c_{23}s_{12}e^{i\delta }& c_{13}c_{23}\end{array}\right)`$ (10) $`\times \left(\begin{array}{ccc}1& 0& 0\\ 0& e^{i\varphi _2}& 0\\ 0& 0& e^{i(\varphi _3+\delta )}\end{array}\right)`$ where $`c_{ij}=\mathrm{cos}\theta _{ij}`$ and $`s_{ij}=\mathrm{sin}\theta _{ij}`$. The extra diagonal phases are present for Majorana neutrinos but do not affect oscillation phenomena. With three neutrinos there are two independent $`\delta m^2`$ and $`\delta m_a^2\delta m_b^2`$ is indicated by the oscillation evidence. The vacuum oscillation probabilities are $$P(\nu _\alpha \nu _\beta )=A_{\alpha \beta }\mathrm{sin}^2\mathrm{\Delta }_aB_{\alpha \beta }\mathrm{sin}^2\mathrm{\Delta }_b+ϵ_{\alpha \beta }J\mathrm{sin}2\mathrm{\Delta }_b,$$ (11) where $`\mathrm{\Delta }_a\delta m_a^2L/4E_\nu `$. $`A_{\alpha \beta }`$ is the amplitude of the leading oscillation, $`B_{\alpha \beta }`$ the amplitude of the sub-leading oscillation and $`J`$ the CP-violating amplitude; all are determined by the $`U`$ matrix elements. The physical variable is $`L/E_\nu `$, where $`L`$ is the baseline from source to detector and $`E_\nu `$ is the neutrino energy. ## IV Matter Effects In matter, $`\nu _e`$ scatter differently from $`\nu _\mu `$ and $`\nu _\tau `$wolf , and the effective neutrino mixing amplitude in matter can be very different from the vacuum amplitudeBPPW ; MS . For the leading oscillation the matter and vacuum oscillation mixings are related in the approximation of constant matter density by $$\mathrm{sin}^22\theta _{13}^m=\frac{\mathrm{sin}^22\theta _{13}}{\left(\mathrm{cos}2\theta _{13}A/\delta m_a^2\right)^2+\mathrm{sin}^22\theta _{13}},$$ (12) where $$A=2\sqrt{2}G_FY_e\rho (x)E_\nu .$$ (13) Here $`Y_e`$ is the electron fraction and $`\rho (x)`$ is the density at path-length $`x`$. The $`\nu _e\nu _\mu `$ (or $`\nu _e\nu _\tau `$) oscillation argument in matter of constant density is $$\mathrm{\Delta }_a^m=\frac{1.27\delta m_a^2(\mathrm{eV}^2)L(\mathrm{km})}{E_\nu (\mathrm{GeV})}\sqrt{\left(\frac{A}{\delta m_a^2}\mathrm{cos}2\theta _{13}\right)^2+\mathrm{sin}^22\theta _{13}}.$$ (14) Resonance enhancements of the oscillation amplitude are possible for $`\delta m_a^2>0`$, while suppression occurs for $`\delta m_a^2<0`$. It is significant that the resonant energies correspond to neutrino energies relevant to the atmospheric and solar anomalies. $`\mathrm{Earth}:`$ $`E_\nu 15\mathrm{GeV}\left({\displaystyle \frac{\delta m_a^2}{3.5\times 10^3\mathrm{eV}^2}}\right)\left({\displaystyle \frac{1.5\mathrm{gm}/\mathrm{cm}^3}{\rho Y_e}}\right)`$ (15) $`\mathrm{Sun}:`$ $`E_\nu 10\mathrm{MeV}\left({\displaystyle \frac{\delta m_b^2}{10^5\mathrm{eV}^2}}\right)\left({\displaystyle \frac{10\mathrm{g}/\mathrm{cm}^3}{\rho Y_e}}\right).`$ (16) ## V Atmospheric Neutrino Oscillations The Kamiokande, SuperKamiokande (SuperK), Macro and Soudan atmospheric neutrino measurementsatmos1 ; atmos2 show a $`\mu /e`$ ratio that is about 0.6 of expectations. The SuperK experimentatmos1 has established the dependence of the $`e`$ and $`\mu `$ event rates on zenith angle, or equivalently the baseline $`L`$ (see Fig. 3). $`N(e)`$ is independent of $`L`$ and validates the $`\nu _e`$ flux calculation (within 20%). $`N(\mu )`$ depletion increases with $`L`$. The muon event distributions are will described by $`\nu _\mu \nu _\tau `$ vacuum oscillations with a $`\nu _\mu `$ survival probability $$P(\nu _\mu \nu _\mu )=1A_{\mu \tau }\mathrm{sin}^2(1.27\delta m_a^2L/E_\nu )$$ (17) with $`\delta m_a^2=3.5\times 10^3\mathrm{eV}^2`$ and maximal or near maximal amplitude $$A_{\mu \tau }=1_{0.2}^{+0.0}(i.e.,|\theta _{32}45^{}|<13^{}).$$ (18) With further data accumulationkajita a slightly lower central value is now indicated ($`\delta m_a^2=2.8\times 10^3\mathrm{eV}^2`$). The $`\nu _\mu \nu _\tau `$ oscillations are not resolved due to smearing of $`L`$ and inferred $`E_\nu `$ values, and equally good fits to the SuperK data are found with oscillation and neutrino decay ($`\nu _2\overline{\nu }_4+J`$) modelsnu-decay , as shown in Fig. 5. A comparison of the unsmeared $`\nu _\mu \nu _\mu `$ probability versus neutrino energy for the oscillation and decay models is shown in Fig. 5. For now we assume the simplest interpretation of the SuperK data, namely oscillations. We note that in the CHOOZ reactor experiment $`\overline{\nu }_e`$ disappearance is not observed at the $`\delta m_a^2`$ scale and the corresponding constraint on 3-neutrino mixing for $`\delta m_a^2=3.5\times 10^3\mathrm{eV}^2`$ is $$A_{\mu e}<0.2,|U_{e3}|<0.23,\theta _{13}<13^{}.$$ (19) ## VI Long-Baseline Experiments Long-baseline experiments are needednu-decay ; bernstein ; BGRW ; shrock ; freund ; rigolin ; camp ; CPV ; derujula ; moreCPV to 1. confirm the atmospheric evidence for $`P(\nu _\mu \nu _\mu )`$ at accelerators; 2. resolve the leading $`\nu _\mu \nu _\mu `$ oscillation and exclude the neutrino decay possibility; 3. precisely measure $`|\delta m_a^2|`$; 4. exclude $`\nu _\mu \nu _s`$ disappearance, although SuperK has now shown that oscillations to sterile neutrinos are excluded at 99% CLkajita ; 5. measure $`|U_{e3}|`$ from $`\nu _e\nu _\mu `$ appearance, which requires a muon decay source for the neutrino beam; 6. determine the sign of $`\delta m_a^2`$ from matter effects in the Earth’s crust; 7. search for CP violationfreund ; rigolin ; camp ; CPV ; derujula ; moreCPV . The first long-baseline experiments will measure the energy dependence of the produced muons and measure the neutral-current to charged-current ratio, to partially address the first four issues listed above. The K2K experiment from KEK to SuperK is in operation, with a baseline $`L=250`$ km and mean neutrino energy $`E_\nu =1.4`$ GeV. The MINOS experiment from Fermilab to Soudan, with $`L=732`$ km and possible energies of $`E_\nu =3,6,12`$ GeV will begin in 2002. A 10% precision on $`|\delta m_a^2|`$ may ultimately be possible at MINOSthesis . The ICANOEicanoe and OPERAopera long-baseline experiments from CERN to Gran Sasso with $`L743`$ km have been approved. ## VII Neutrino Factories Muon storage rings could provide intense neutrino beams ($`10^{19}\text{}10^{21}`$ per year) that would yield thousands of charged-current neutrino interactions in a reasonably sized detector (10–50 kt) anywhere on Earthnufactories . These neutrino factories would have pure neutrino beams ($`\nu _e,\overline{\nu }_\mu `$ from stored $`\mu ^+`$ and $`\overline{\nu }_e,\nu _\mu `$ from stored $`\mu ^{}`$) with 50% $`\nu _e`$ or $`\overline{\nu }_e`$ components. Detection of wrong-sign muons (the muons with opposite sign to the charge current from the beam muon neutrino) would signal $`\nu _e\nu _\mu `$ or $`\overline{\nu }_e\overline{\nu }_\mu `$ appearance oscillations. We now discuss the capability of a neutrino factory with $`2\times 10^{20}`$ muons a year and a 10 kt detector to resolve the issues raised in the preceding section. Figure 6 shows the precision attainable in $`\delta m_{32}^2`$ and $`\mathrm{sin}^22\theta _{23}`$ parameters through $`\nu _\mu `$ survival measurements based on an $`E_\mu =30`$ GeV storage ring and a baseline of $`L=2800`$ km. A statistical precision of a few % on $`\mathrm{sin}^22\theta _{23}`$ is possible in $`\nu _\mu `$ disappearance measurements. This accuracy in measuring $`\mathrm{sin}^22\theta _{23}`$ would differentiate the bimaximal model predictionbimax1 ; bimax2 of $`\mathrm{sin}^22\theta _{23}=1`$ from the democratic model predictiondemoc of $`\mathrm{sin}^22\theta _{23}=8/9`$. Figure 7 gives the tau-lepton yields at $`E_\mu =10`$, 30 and 50 GeV for baselines of $`L=732`$, 2800, and 7332 km in a 1 kt detector for an intensity of $`2\times 10^{20}`$ neutrinos. There would be hundreds of events per year from $`\nu _\mu \nu _\tau `$ oscillations; however, the signal of $`\nu _e\nu _\tau `$ would be difficult with the $`2\times 10^{20}`$ luminosity. The wrong-sign muon event rates are approximately proportional to $`\mathrm{sin}^22\theta _{13}`$. Figure 10 shows the predicted numbers of eventsBGRW with $`\mathrm{sin}^22\theta _{13}=0.04`$ for a 10 kt detector at $`L=2900`$ km with an intensity of $`2\times 10^{20}`$ neutrinos. The observation of $`\nu _e\nu _\mu `$ and $`\overline{\nu }_e\overline{\nu }_\mu `$ appearance oscillations at baselines long enough to have significant matter effects will allow a determination of the sign of $`\delta m_{32}^2`$, and thus determine the pattern of the masses; see Fig. 10. A proof of the principle that the sign of $`\delta m^2`$ can be so determinedBGRW is given in Fig. 10 for a baseline $`L=2800`$ km. In $`\mu ^+`$ appearance, $`\delta m_{32}^2>0`$ gives a smaller rate and harder spectrum than $`\delta m_{32}^2<0`$, while the results are opposite in $`\mu ^{}`$ appearance. In optimizing $`E_\mu `$ and $`L`$ for long-baseline experiments to find the sign of $`\delta m_{32}^2`$, $`L=732`$ km is too short (matter effects are small) and $`L=7332`$ km is too far (event rates are low)BGRW ; rigolin . The sensitivity to determine the sign of $`\delta m_{32}^2`$ improves linearly with $`E_\mu `$. There is a tradeoff between energy, detector size and muon beam intensityBGRW . The $`\nu _e\nu _\mu `$ sensitivity on $`\mathrm{sin}^22\theta _{23}`$ and sign $`\delta m_{32}^2`$ for a 10 kt detector is shownBGRW in Fig. 11 for $`E_\mu =10`$, 30, and 50 GeV. ## VIII Solar Neutrino Oscillations The solar neutrino experimentssolarexpts sample different $`\nu _e`$ energy ranges and find different flux deficits compared to the Standard Solar Model (SSM)bahcall-sm as follows: $$\begin{array}{ccc}\nu _e{}_{}{}^{71}\mathrm{Ga}{}_{}{}^{71}\mathrm{Ge}e\hfill & \begin{array}{c}\mathrm{GALLEX}\hfill \\ \mathrm{SAGE}\hfill \end{array}\hfill & \begin{array}{c}0.60\pm 0.06\hfill \\ 0.52\pm 0.06\hfill \end{array}\hfill \\ \nu _e{}_{}{}^{37}\mathrm{Cl}{}_{}{}^{37}\mathrm{Ar}e\hfill & \mathrm{Homestake}\hfill & 0.33\pm 0.03\hfill \\ \nu e\nu e\hfill & \mathrm{SuperK}\hfill & 0.47\pm 0.02\hfill \end{array}$$ (20) Thus the $`\nu _e`$ survival probability is inferred to be energy dependent. Global oscillation fitssolardata have been made using floating <sup>8</sup>B and hep flux normalizations which are somewhat uncertain in the SSM. The relative normalizations from the fits range from 0.5 to 1.2 for the <sup>8</sup>B flux and 1 to 25 for the hep flux. These global fits include the data on 1. total rates, assuming all the experiments are okay; the different Cl suppression ratio plays a vital role; 2. the night-day asymmetry, which is observed at the 2$`\sigma `$ levelkajita ; the large angle matter solution gives night rates $`>`$ day rates; 3. seasonal dependence beyond $`1/r^2`$, which can occur for vacuum solutions. The oscillation analysessolardata generally agree on the allowed $`\delta m_{21}^2`$ and $`\mathrm{sin}^22\theta _{12}`$ regions for acceptable solutions. Typical candidate solar solutions are given in Table 1. In the case of vacuum oscillations (VO) several discrete regions of $`\delta m_{21}^2`$ are possible. ## IX 3-Neutrino Mixing Matrix Once the solar oscillation solution is pinned down, and $`\theta _{12}`$ is thus determined, we will have approximate knowledge of the mixing angles of the 3-neutrino matrix, with $`\theta _{23}\pi /4`$ and $`\theta _{13}0`$ from the atmospheric and CHOOZ data. Upcoming experiments are expected to shed light on the solar solution. In the SNO experimentSNO , which is now taking data, and the upcoming ICARUS experimenticarus , the high energy $`\nu _e`$ CC events may distinguish LAM, SAM, and LOW solutions with large hep flux contributions from the VO or the SAM sterile neutrino solutionsBKS99 . Also, the neutral-current to charged-current ratio will distinguish active from sterile oscillations. The Borexinoborex experiment can measure the VO seasonal variation of the <sup>7</sup>Be line flux7Be . The KamLand reactor experimentkamland to measure the $`\overline{\nu }_e`$ survival probability will be sensitive to the LAM and LOW solar solutionsmuray-kam . The CP phase $`\delta `$ may be measurable at a neutrino factoryfreund ; rigolin ; camp ; derujula if the solar solution is LAM. The CP violation comes in only at the sub-leading oscillation scale. The CP and T asymmetries for $`\nu _\mu \nu _e`$ oscillations are $`A_{\alpha \beta }^{CP}`$ $`=`$ $`{\displaystyle \frac{P(\nu _\alpha \nu _\beta )P(\overline{\nu }_\alpha \overline{\nu }_\beta )}{P(\nu _\alpha \nu _\beta )+P(\overline{\nu }_\alpha \overline{\nu }_\beta )}},`$ (21) $`A_{\alpha \beta }^T`$ $`=`$ $`{\displaystyle \frac{P(\nu _\alpha \nu _\beta )P(\nu _\beta \nu _\alpha )}{P(\nu _\alpha \nu _\beta )+P(\nu _\beta \nu _\alpha )}}.`$ (22) An apparent CP-odd asymmetry is induced by matter. ## X Models For maximal mixing in both atmospheric and solar sectors, there is an unique mixing matrixbimax1 $$U=\left(\begin{array}{ccc}1/\sqrt{2}& 1/\sqrt{2}& 0\\ 1/2& 1/2& 1/\sqrt{2}\\ 1/2& 1/2& 1/\sqrt{2}\end{array}\right).$$ (23) In this bimaximal mixing model, there would be no CP-violating effects. However, because $`U_{e3}=0`$, long-baseline experiments would have some sensitivity to the sub-leading LAM solar scale oscillations. Many unification modelsRMreview ; king predict that the neutrino masses are Majorana and hierarchical, there is no cosmologically significant dark matter, and the SAM solar solution (small $`\theta _{12}`$ mixing) obtains. ## XI Absolute Neutrino Masses Oscillation phenomena determine only mass-squared differences, leaving the absolute mass scale unknown. However, because the atmospheric and solar $`\delta m^2`$ values are $`(1\mathrm{eV})^2`$, all mass eigenvalues are approximately degenerate if at the $`1`$ eV scale. Thus all neutrino mass eigenvalues are boundedBWW by the tritium limit from the Troitsk and Mainz experiments, $$m_j<3\mathrm{eV}\mathrm{for}j=1,2,3.$$ (24) Neutrinoless double-beta decay (0$`\nu \beta \beta `$) provides a probe of Majorana neutrino mass0nbb via the diagram in Fig. 12. The rate is proportional to the $`\nu _e\nu _e`$ element of the neutrino mass matrix. The present limit from the Heidelberg experimentbaudis is $$M_{\nu _e\nu _e}<0.2\mathrm{eV}\times f,$$ (25) where the factor $`f`$ represents the theoretical uncertainty in the nuclear matrix elements, which might be as large as a factor of 3. The $`0\nu \beta \beta `$ limit translates to a bound on the summed neutrino Majorana masses ofBW99 $$m_\nu <0.75\mathrm{eV}\times f$$ (26) in the SAM solar solution. No similar constraints apply to the LAM, LOW or VO solutions where the bound can be satisfied by having opposite CP parity of $`\nu _1`$ and $`\nu _2`$ mass eigenstates. Future sensitivity down to $$|M_{\nu _e\nu _e}|=0.01\mathrm{eV}$$ (27) is expectedfuture , which would provide sensitivity down to $$m_\nu =0.08\mathrm{eV}\times f$$ (28) in the SAM solution. ## XII Beyond 3 Neutrinos The LSND evidence for $`\nu _\mu \nu _e`$ oscillations with $`\delta m^21\mathrm{eV}^2`$, $`\mathrm{sin}^22\theta 10^2`$ requires a $`\delta m^2`$ scale distinct from the atmospheric and solar oscillation scales, and thus a sterile neutrino state would be needed to explain all the oscillation phenomena. Then to also satisfy limits from CDHS acceleratorcdhs and Bugey reactorbugey experiments, the mass hierarchy must be two separated pairsdoublets , as shown in Fig. 13. Such a scenario would allow even more interesting effects at a neutrino factory, such as large CP violationhattori , since both the leading and sub-leading oscillation scales would be accessible. The MiniBooNE experimentminiboone will settle whether the LSND evidence is real. Other interest in sterile neutrinos comes from $`r`$-process nucleosynthesis if it occurs in supernovaebaha . ## XIII Neutrino Mass from Cosmology Measurements of the power spectrum by the MAP and PLANCK satellites may determine $`m_\nu `$ down to $`0.4`$ eVcosmo . The heights of the acoustic peaks can also decide how the mass is distributed among the neutrino flavorsfalk , as illustrated in Fig. 14. ## XIV Summary We have entered an exciting new era in the study of neutrino masses and mixing. From the SuperK evidence on atmospheric neutrino oscillations, we already have a surprising amount of information about the neutrino mixing matrix (near maximal $`\mathrm{sin}^22\theta _{23}`$ and near minimal $`\mathrm{sin}^22\theta _{13}`$). The SuperK, SNO, Borexino, KamLand, and ICARUS experiments are expected to differentiate among the candidate solar oscillation possibilities and determine $`\mathrm{sin}^22\theta _{12}`$. MiniBooNE will tell us whether a sterile neutrino is mandated. Neutrino factories will study the leading oscillations, determine the sign of $`\delta m_a^2`$, measure $`U_{e3}`$, and possibly detect CP violation. The GENIUS $`0\nu \beta \beta `$ experiment and the MAP and PLANCK satellite measurements of the power spectrum will probe the absolute scale of neutrino masses. There is a synergy of particle, physics, nuclear physics, and cosmology occurring in establishing the fundamental properties of neutrinos. A theoretical synthesis should emerge from these experimental pillars. ## Acknowledgments I thank my collaborators K. Whisnant, T. Weiler, S. Pakvasa, S. Geer, R. Raja, J. Learned, P. Lipari, and M. Lusignoli. This research was supported in part by the U.S. Department of Energy under Grant No. DE-FG02-95ER40896 and in part by the University of Wisconsin Research Committee with funds granted by the Wisconsin Alumni Research Foundation.
warning/0003/math0003054.html
ar5iv
text
# Projectively Quantization Map ## 1 Introduction Quantization procedure proposed in this paper is deals with the space of linear differential operators and the corresponding space of symbols viewed as modules over the group of diffeomorphisms $`\mathrm{Diff}(M)`$ and the Lie algebra of vectors fields $`\mathrm{Vect}(M).`$ This method of quantization have been introduced in the recent papers (, , ). Let $`𝒟_{\lambda ,\mu }(M)`$ be the space of linear differential operators from the space of $`\lambda `$densities with the space of $`\mu `$densities. The corresponding space of symbols, $`\mathrm{Pol}_\delta (T^{}M),`$ is the space of polynomials on $`T^{}M`$ with values in the space of $`\delta `$densities, where $`\delta =\mu \lambda .`$ We call quantization map, a linear map $$Q_{\lambda ,\mu }:\mathrm{Pol}_\delta (M)𝒟_{\lambda ,\mu }(M),$$ (1.1) that is bijective and preserves the principal symbol (see , , ). There is no quantization map (1.1) equivariant with respect to the action of the group $`\mathrm{Diff}(M)`$. It is natural to consider a subgroup $`G\mathrm{Diff}(M)`$ (of finite dimension) and to restrict the action of $`\mathrm{Diff}(M)`$ on the subgroup $`G.`$ There are two interesting cases: If $`M=𝐑^n`$ is endowed with a flat projective structure, the quantization map is given in . This map is equivariant with respect to the action of the group of projective transformations $`\mathrm{SL}_{n+1}\mathrm{Diff}(𝐑^n)`$. If $`M=𝐑^n`$ is endowed with a flat conformal structure, the quantization map is given in , it is equivariant with respect to the action of the group of conformal diffeomorphisms $`SO(p+1,q+1)\mathrm{Diff}(𝐑^n),`$ where $`(p+q=n).`$ (See also , , , for the one dimensional case.) A natural and well-known way to define a quantization map is to fix an affine connection on $`M`$ (see, e.g. ). However, there is no canonical quantization map associated to a given connection. The purpose of this paper is to study the quantization map (1.1) between the space of second-order symbols and the space of second-order linear differential operators satisfying the following properties: 1. It is projectively invariant, i.e. it depend only on the projective class of the affine connection $`\mathrm{\Gamma }.`$ 2. If $`M=𝐑^n`$ with a flat projective structure, this isomorphism is equivariant with respect to the action of the group of projective transformations $`\mathrm{SL}_{n+1}`$ (resp. infinitesimal projective transformations $`\mathrm{sl}_{n+1}`$). The method used in this paper follows that of the recent preprint . ## 2 Space of linear differential operators Let $`M`$ be a manifold of dimension $`n`$ endowed with an affine connection $`\mathrm{\Gamma }.`$ We are interested in defining a two parameter family of $`\mathrm{Diff}(M)`$module (resp. $`\mathrm{Vect}(M)`$module) on the space of linear differential operators. This space was recently studied in recent papers (, , , , , , , ). ### 2.1 Space of tensor densities For simplicity, we assume $`M`$ oriented throughout this paper. The space of tensor densities on $`M,`$ $`_\lambda (M),`$ or $`_\lambda `$ for simplify, is the space of sections of the line bundle $`(\mathrm{\Lambda }^nT^{}M)^\lambda ,`$ where $`\lambda 𝐂.`$ As a vector space, tensor densities are isomorphic to the space of complexified functions, but the structure of $`\mathrm{Diff}(M)`$-module is different. Let us explicit this action: Let $`f\mathrm{Diff}(M)`$ and $`\varphi _\lambda .`$ In a local coordinates $`(x^i)`$, the action is given by $$f^{}\varphi =\varphi f^1(J_{f^1})^\lambda ,$$ (2.2) where $`J_f=\left|\frac{Df}{Dx}\right|`$ is the Jacobian of $`f`$. In the case $`\lambda =0,`$ $`1`$, the action (2.2) is precisely the standard action of $`\mathrm{Diff}(M)`$ on the space of functions and differential forms of degree $`n`$ respectively. Differentiating the action of the flow of a vector field, one gets the corresponding representation of $`\mathrm{Vect}(M).`$ $$L_X^\lambda (\varphi )=X^i_i(\varphi )+\lambda _i(X^i)\varphi ,$$ (2.3) where $`X=X^i_i\mathrm{Vect}(M).`$ The formulæ (2.2), (2.3) do not depend on the choice of coordinates. Let us now recall the definition of covariant derivative on tensor densities (cf. ). Let $``$ be the covariant derivative associated to the affine connection $`\mathrm{\Gamma }.`$ If $`\varphi _\lambda `$, then $`\varphi \mathrm{\Omega }^1(M)_\lambda ,`$ given, in a local coordinates, by the formula: $$_i\varphi =_i\varphi \lambda \mathrm{\Gamma }_i\varphi ,$$ (2.4) with $`\mathrm{\Gamma }_i=\mathrm{\Gamma }_{il}^l`$ (summation is understood on repeated indices). ### 2.2 Space of linear differential operators Consider the space of linear differential operators acting on tensor densities $$A:_\lambda _\mu .$$ The action of $`\mathrm{Diff}(M)`$ on $`𝒟(M)`$ depends on two parameters $`\lambda `$ and $`\mu `$. This action is given by the equation: $$f_{\lambda ,\mu }(A)=f^{}Af_{}^{}{}_{}{}^{1},$$ (2.5) where $`f^{}`$ is the action (2.2) of $`\mathrm{Diff}(M)`$ on $`_\lambda `$. Differentiating the action of the flow of a vector field, one gets the corresponding representation of $`\mathrm{Vect}(M).`$ $`L_X^{\lambda ,\mu }(A)=L_X^\mu AAL_X^\lambda ,`$ (2.6) where $`X\mathrm{Vect}(M).`$ These formulæ do not depend on the choice of a system of coordinates. Notation. Denote $`𝒟^k(M)`$ the space of $`k`$order linear differential operators. In a local coordinates $`(x^i),`$ one can write $`A𝒟^k(M)`$ $$A=a_k^{i_1,\mathrm{},i_k}\frac{}{x^{i_1}}\mathrm{}\frac{}{x^{i_k}}+\mathrm{}+a_1^i\frac{}{x^i}+a_0,$$ with the coefficients $`a_k^{i_1,\mathrm{},i_k}=a_k^{i_1,\mathrm{},i_k}(x^1,\mathrm{},x^n)C^{\mathrm{}}(M).`$ We have then a filtration $$𝒟^0𝒟^1\mathrm{}𝒟^k\mathrm{}$$ Denote by $`𝒟_{\lambda ,\mu }`$ the module of linear differential operators on $`M`$ endowed with the action of $`\mathrm{Diff}(M)`$ (resp. $`\mathrm{Vect}(M)`$) given by (2.5) (resp. (2.6)). The space of $`k`$order linear differential operators, denoted by $`𝒟_{\lambda ,\mu }^k,`$ is a $`\mathrm{Diff}(M)`$-submodule (resp. $`\mathrm{Vect}(M)`$submodule) of $`𝒟_{\lambda ,\mu }.`$ ###### Remark 2.1 The space of linear differential operators viewed as a module over the group of diffeomorphisms is a classical object (see e.g. ). For example, in the case $`M=S^1,`$ the space of sturm Liouville operators $`\frac{d^2}{dx^2}+u(x)`$ is viewed as a submodule of $`𝒟_{\frac{1}{2},\frac{3}{2}}^2.`$ Also, the modules $`𝒟_{\frac{1k}{2},\frac{1+k}{2}}^k`$ was considered in . ## 3 Space of symbols The space of symbols, $`\mathrm{Pol}(T^{}M),`$ is the space of functions on the cotangent bundle $`T^{}M`$ polynomial on the fibers. In a local coordinate system $`(x_i,\xi _i),`$ one can write $$T=\underset{l=0}{\overset{k}{}}T^{i_1,\mathrm{},i_l}\xi _{i_1}\mathrm{}\xi _{i_l},$$ with $`T^{i_1,\mathrm{},i_l}(x^1,\mathrm{},x^n)C^{\mathrm{}}(M).`$ One defines a one parameter family of $`\mathrm{Diff}(M)`$module (resp. $`\mathrm{Vect}(M)`$module) on the space of symbols by $$\mathrm{Pol}_\delta (T^{}M):=\mathrm{Pol}(T^{}M)_\delta .$$ Let us explicit this action. Take $`f\mathrm{Diff}(M)`$ and $`X\mathrm{Vect}(M)`$. Then, in a local coordinates $`(x_i,\xi _i),`$ one has: $`f_\delta (T)`$ $`=`$ $`f_{}T(J_{f^1})^\delta ,`$ (3.7) $`L_X^\delta (T)`$ $`=`$ $`L_X(T)+\delta \mathrm{D}(X)T,`$ (3.8) where $$L_X=X^i_i\xi _j_i(X^j)_{\xi _i},\mathrm{D}(X)=_iX^i.$$ The space of symbols admits a graduation $$\mathrm{Pol}_\delta (T^{}M)=\underset{k=0}{\overset{\mathrm{}}{}}\mathrm{Pol}_{\delta ,k}(T^{}M),$$ where $`\mathrm{Pol}(T^{}M)_{\delta ,k}`$ are the homogeneous polynomials of degree $`k`$ on $`T^{}(M).`$ This graduation is $`\mathrm{Diff}(M)`$invariant. Throughout this paper, we will identify the space of symbols with the space of symmetric contravariant tensor fields on $`M.`$ ## 4 Flat projective structure and projectively equivalent connection Let $`M`$ be a manifold of dimension $`n.`$ Recall two notions on projective geometry, the notion of flat projective structure and the notion of projectively equivalent connections (see ). ### 4.1 Flat projective structure A manifold $`M`$ admits a flat projective structure if there exists an atlas $`\{\varphi _i\}`$ such that the local transformations $`\varphi _i\varphi _j^1`$ are projective transformations. The most interesting case is when $`M=𝐑^n.`$ In this case, the group $`\mathrm{SL}_{n+1}`$ acts locally on $`𝐑^n`$ by projective transformations. Choosing a local coordinate system, the Lie algebra $`\mathrm{sl}_{n+1}`$ can be identified with the subalgebra of $`\mathrm{Vect}(𝐑^n)`$ generated by the vector fields: $$_i,x^i_j,x^ix^j_j.$$ The projective Lie algebra $`\mathrm{sl}_{n+1}`$ is a maximal subalgebra of the Lie algebra of polynomial vector fields on $`𝐑^n`$ (cf. ). ### 4.2 Projectively equivalent connections The notion of projectively equivalent connection is an old notion related to projective geometry of the “paths” studied by H. Weyl in and T.Y. Thomas in . Weyl gives the following definition: Two affine connection without torsion, with Christoffel symbols $`\stackrel{~}{\mathrm{\Gamma }}_{jk}^i`$ and $`\mathrm{\Gamma }_{jk}^i`$ given on the same system of coordinate $`x^1,\mathrm{},x^n`$, are projectively equivalent, if there exists a differential 1-form with components $`\omega _i,`$ such that $$\stackrel{~}{\mathrm{\Gamma }}_{jk}^i=\mathrm{\Gamma }_{jk}^i+\delta _j^i\omega _k+\delta _k^i\omega _j.$$ (4.9) Geometrically, two affine connections without torsion projectively equivalent give the same unparameterized geodesics (cf. , ). An affine connection $`\mathrm{\Gamma }`$ is said to be projectively flat, if they can be written: $$\mathrm{\Gamma }_{jk}^i=\frac{1}{n+1}\left(\delta _j^i\mathrm{\Gamma }_k+\delta _k^i\mathrm{\Gamma }_j\right).$$ (4.10) A manifold $`M`$ endowed with an affine connection $`\mathrm{\Gamma }`$, admits a flat projective structure if and only if the connection $`\mathrm{\Gamma }`$ is projectively flat (cf. ). ## 5 Main theorems In this section, we will give the quantization map between the space of second-order symbols and the space of second-order linear differential operators. First, decompose the space of symbols into a direct sum $$\mathrm{Pol}_{\delta ,2}(T^{}M)\mathrm{Pol}_{\delta ,1}(T^{}M),$$ where $`\mathrm{Pol}_{\delta ,2}(T^{}M)`$ is the space of symbols of degree 2, and $`\mathrm{Pol}_{\delta ,1}(T^{}M)`$ the space of symbols of degree less or equal 1. We will construct a quantization map on each of these spaces. In the case of first order symbols, there exists a quantization map that commutes with the action of $`\mathrm{Diff}(M)`$ and $`\mathrm{Vect}(M)`$ (see , ). ###### Theorem 5.1 For any $`\delta 1,`$ the map $`Q_{\lambda ,\mu }^1:\mathrm{Pol}_{\delta ,1}(M)𝒟_{\lambda ,\mu }^1(M)`$ given by $$Q_{\lambda ,\mu }^1(T)=T^i_i+\alpha _i(T^i)+T^0$$ (5.11) where $`T=T^i\xi _i+T^0,`$ and $$\alpha =\frac{\lambda }{1\delta }$$ (5.12) is a projectively invariant isomorphism (i.e. it depends only on the projective class of the affine connection $`\mathrm{\Gamma }.`$) Proof of Theorem 5.1. Let $`\stackrel{~}{\mathrm{\Gamma }}`$ be a symmetric affine connection projectively equivalent to $`\mathrm{\Gamma }.`$ Denote by $`\stackrel{~}{Q}_{\lambda ,\mu }^1`$ the quantization map written with the connection $`\stackrel{~}{\mathrm{\Gamma }}`$. We must show $$\stackrel{~}{Q}_{\lambda ,\mu }^1=Q_{\lambda ,\mu }^1.$$ We need some formulæ (see ): Recall the covariant derivative on the space of 1-order contravariant tensor fields: Let $`T^i`$ be a tensor, then one has: $`_j(T^i)`$ $`=`$ $`_jT^i+\mathrm{\Gamma }_{jl}^iT^l\delta \mathrm{\Gamma }_jT^i.`$ (5.13) Using the formulæ (4.9), (5.13) one obtains $$\stackrel{~}{}_i(\varphi )=_i(\varphi )\lambda (n+1)\omega _i\varphi ,\stackrel{~}{}_i(T^i)=_i(T^i)+(1\delta )(1+n)\omega _iT^i.$$ Then after a straightforward calculation: $$\stackrel{~}{Q}_{\lambda ,\mu }^1(T)=Q_{\lambda ,\mu }^1(T)+(1+n)(\alpha (1\delta )\lambda )\omega _iT^i.$$ Hence $`\stackrel{~}{Q}_{\lambda ,\mu }^1=Q_{\lambda ,\mu }^1`$ if and only if $`\alpha `$ is given as in (5.12). ###### Remark 5.2 1. If $`M`$ is endowed with a flat projective structure, the isomorphism (5.11) is the unique provided it preserves the principal symbol (cf. ). 2. In the case $`\delta =1,`$ the modules are still isomorphic if $`(\lambda ,\mu )=(0,1).`$ This isomorphism is given by (5.11) with an arbitrary $`\alpha =0.`$ Let us give the quantization map on the space of homogeneous symbols of degree $`2.`$ ###### Theorem 5.3 If $`n2,`$ for any $`\delta \frac{n+3}{n+1},\frac{n+2}{n+1},`$ there exists a projectively invariant isomorphism $`Q_{\lambda ,\mu }^2:\mathrm{Pol}_{\delta ,2}(T^{}M)𝒟_{\lambda ,\mu }^2(M)`$ given by $$Q_{\lambda ,\mu }^2(T)=T^{ij}_i_j+\beta _1_jT^{ij}_i+\beta _2_i_j(T^{ij})+\beta _3R_{ij}T^{ij},$$ (5.14) where $`T(\xi )=T^{ij}\xi _i\xi _j,`$ the coefficients $`\beta _1,\beta _2,\beta _3`$ are as follows $`\beta _1`$ $`=`$ $`{\displaystyle \frac{2+2\lambda (n+1)}{2+(1+n)(1\delta )}}`$ $`\beta _2`$ $`=`$ $`{\displaystyle \frac{\lambda (n+1)(1+\lambda (n+1))}{((1\delta )(1+n)+1)((1\delta )(1+n)+2)}}`$ (5.15) $`\beta _3`$ $`=`$ $`{\displaystyle \frac{\lambda (\mu 1)(n+1)^2}{(1n)((1\delta )(1+n)+1)}}`$ and $`R_{ij}`$ denote the components of Ricci tensor of the connection $`\mathrm{\Gamma }.`$ ###### Corollary 5.4 If $`M`$ is endowed with a flat projective structure then: 1. The isomorphism (5.14) has the following form: $`Q_{\lambda ,\mu }(T)`$ $`=`$ $`T^{ij}_i_j+\beta _1_jT^{ij}_i+\beta _2_i_jT^{ij},`$ (5.16) where the constants $`\beta _1,\beta _2`$ are as in (5.3). 2. It is the unique map equivariant with respect to the action of $`\mathrm{SL}_{n+1}`$ (resp. $`\mathrm{sl}_{n+1}`$) that preserves the principal symbols (cf. ). Proof of the Theorem 5.3 Let $`\stackrel{~}{\mathrm{\Gamma }}`$ be a connection projectively equivalent to $`\mathrm{\Gamma }.`$ Denote by $`\stackrel{~}{Q}_{\lambda ,\mu }^2`$ the quantization map written with $`\stackrel{~}{\mathrm{\Gamma }}.`$ We need some formulæ (see ): The covariant derivative on the space of 2-order contravariant tensor fields reads: $`_k(T^{ij})`$ $`=`$ $`_kT^{ij}+\mathrm{\Gamma }_{lk}^iT^{lj}+\mathrm{\Gamma }_{lk}^jT^{il}\delta \mathrm{\Gamma }_kT^{ij}.`$ (5.17) The second-order term in $`Q_{\lambda ,\mu }^2`$ reads: $`T^{ij}_i_j`$ $`=`$ $`T^{ij}_i_j+(T^{jk}\mathrm{\Gamma }_{jk}^i+2\lambda T^{ij}\mathrm{\Gamma }_j)_i`$ $`+T^{ij}(\lambda ^2\mathrm{\Gamma }_i\mathrm{\Gamma }_j+\lambda _i\mathrm{\Gamma }_j),`$ the first-order term in $`Q_{\lambda ,\mu }^2`$ reads: $`_jT^{ij}_i`$ $`=`$ $`_jT^{ij}_i(T^{jk}\mathrm{\Gamma }_{jk}^i+(1\delta )T^{ij}\mathrm{\Gamma }_j)_i`$ $`+\lambda (_iT^{ij})\mathrm{\Gamma }_j\lambda T^{ij}(\mathrm{\Gamma }_{ij}^k\mathrm{\Gamma }_k+(1\delta )\mathrm{\Gamma }_i\mathrm{\Gamma }_j),`$ and the zero-order part of $`Q_{\lambda ,\mu }^2`$ reads: $`_j_jT^{ij}`$ $`=`$ $`_i_jT^{ij}2(1\delta )(_iT^{ij})\mathrm{\Gamma }_j(_iT^{jk})\mathrm{\Gamma }_{jk}^i`$ $`T^{ij}(_k\mathrm{\Gamma }_{ij}^k+(1\delta )_i\mathrm{\Gamma }_j2\mathrm{\Gamma }_{ik}^l\mathrm{\Gamma }_{jl}^k(12\delta )\mathrm{\Gamma }_{ij}^k\mathrm{\Gamma }_k(1\delta )^2\mathrm{\Gamma }_i\mathrm{\Gamma }_j).`$ Now after calculation one has: $`\stackrel{~}{Q}_{\lambda ,\mu }(T)=Q_{\lambda ,\mu }(T)+[2\beta _2+(1+n)(\lambda \beta _1+2\eta _\delta \beta _2)]_iT^{ij}\omega _j`$ $`+[2\beta _12+(1+n)(2\lambda +\eta _\delta \beta _1)]T^{ij}\omega _j_i`$ $`+[(1+n)(\lambda +\eta _\delta \beta _2)+2\beta _2+(1n)\beta _3]T^{ij}_i\omega _j`$ $`+[(1+n)(\lambda \eta _\delta \beta _2)2\beta _2+\beta _3(n1)]T^{jk}\mathrm{\Gamma }_{jk}^i\omega _i`$ $`+[(1+n)^2(\lambda ^2+\eta _\delta (\delta \beta _2\lambda \beta _1))+2(1+n)(\lambda (1\beta _1)+\delta \beta _2)+(n1)\beta _3]T^{ij}\omega _i\omega _j`$ where $`\eta _\delta =1\delta .`$ Hence, $`\stackrel{~}{Q}_{\lambda ,\mu }^2=Q_{\lambda ,\mu }^2`$ if and only if the constants $`\beta _1,\beta _2,\beta _3`$ are given as in (5.3). Proof of the Corollary 5.4. In this case (see section 4.2), the connection $`\mathrm{\Gamma }`$ can be written, in the coordinates of the flat projective structure, in the form $$\mathrm{\Gamma }_{ij}^k=\frac{1}{n+1}(\delta _i^k\mathrm{\Gamma }_j+\delta _j^k\mathrm{\Gamma }_i).$$ Substituting this formula to the equations (5), (5), (5), and, finally to the map (5.4) one gets the expression (5.16). The proof of the part 2) is given in . Let us study the particular values of $`\delta `$ called “resonant”: ###### Proposition 5.5 In the resonant case $`\delta =\frac{n+2}{n+1},\frac{n+3}{n+1}`$, the modules are still isomorphic with the particular values of $`\lambda ,\mu ,\beta _1,\beta _2,\beta _3,`$ given in the table I bellow. Proof of the proposition 5.5. Replace the particular values of $`\delta `$ in the formula (5). hence, $`Q_{\lambda ,\mu }^2=\stackrel{~}{Q}_{\lambda ,\mu }^2,`$ if and only if the constants $`\lambda ,\mu ,\beta _1,\beta _2,\beta _3`$ is given as in the table I. ###### Remark 5.6 In contrast with the non-resonant case, if $`M`$ is flat and $`\delta =\frac{n+3}{n+1},`$ the isomorphism is not unique. There is a family of isomorphisms with arbitrary constant $`\beta _2.`$ | $`\delta `$ | $`\lambda `$ | $`\mu `$ | $`\beta _1`$ | $`\beta _2`$ | $`\beta _3`$ | | --- | --- | --- | --- | --- | --- | | $`{\displaystyle \frac{n+3}{n+1}}`$ | $`{\displaystyle \frac{1}{n+1}}`$ | $`{\displaystyle \frac{n+2}{n+1}}`$ | $`2\beta _2`$ | . | $`{\displaystyle \frac{1}{1n}}`$ | | $`{\displaystyle \frac{n+2}{n+1}}`$ | $`0`$ | $`{\displaystyle \frac{n+2}{n+1}}`$ | $`2`$ | $`0`$ | $`0`$ | | $`{\displaystyle \frac{n+2}{n+1}}`$ | $`{\displaystyle \frac{1}{n+1}}`$ | $`1`$ | $`0`$ | $`0`$ | $`{\displaystyle \frac{1}{1n}}`$ | Table I. It would be interesting to obtain an analogue of the formula (5.14) in the case of higher-order differential operators. Acknowledgments. I am embedded to C. Duval and V. Ovsienko for the statement of the problem and numerous fruitful discussions.
warning/0003/nucl-th0003062.html
ar5iv
text
# FZJ-IKP(TH)-2000-09 The pion charge radius from charged pion electroproduction \[ ## Abstract We analyze a low–energy theorem of threshold pion electroproduction which allows one to determine the charge radius of the pion. We show that at the same order where the radius appears, pion loops induce a correction to the momentum dependence of the longitudinal dipole amplitude $`L_{0+}^{()}`$. This model–independent correction amounts to an increase of the pion charge radius squared from the electroproduction data by about 0.26 fm<sup>2</sup>. It sheds light on the apparent discrepancy between the recent determination of the pion radius from electroproduction data and the one based on pion–electron scattering. PACS nos.: 25.30.Rw, 14.40.Aq, 12.39.Fe Keywords: Pion electroproduction, pion charge radius, chiral perturbation theory \] The charge (vector) radius of the pion is a fundamental quantity in hadron physics. It can essentially be determined in two ways. One method is pion scattering off electrons (or electron–positron annihilation into pion pairs), this leads a pion root–mean–square (rms) radius of $$r_\pi ^2_V^{1/2}=(0.663\pm 0.006)\mathrm{fm},$$ (1) if one insists on the correct normalization of the pion charge (vector) form factor, $`F_\pi ^V(0)=1`$ (in units of the elementary charge $`e`$). A more recent determination of the pion vector radius from low–momentum space– and time–like form factor data based on a very precise two–loop chiral perturbation theory representation gives $$r_\pi ^2_V^{1/2}=(0.661\pm 0.012)\mathrm{fm},$$ (2) consistent with the value given above. This number can be semi–quantitatively be understood in a naive vector meson dominance picture, $`r_\pi ^2_V^{1/2}=\sqrt{6}/M_\rho 0.63`$fm. The second method is based on charged pion electroproduction, $`\gamma ^{}p\pi ^+n`$. Here, $`\gamma ^{},p,n`$ and $`\pi ^+`$ denote the virtual photon, the proton, the neutron and the positively charged pion, in order. The unpolarized cross section in parallel kinematics decomposes into a transversal and a longitudinal piece.<sup>*</sup><sup>*</sup>*For a textbook discussion on this issue, we refer the reader to Ref.. While the former is sensitive to the the nucleon axial radius, the latter is quite sensitive to the pion form factor, i.e. to the pion radius for small momentum transfer. A recent measurement at the Mainz Microtron MAMI–II led to a pion radius of $$r_\pi ^2_V^{1/2}=(0.74\pm 0.03)\mathrm{fm},$$ (3) which is a sizeably larger value than the one obtained from $`\pi e`$ scattering. It was hinted in Ref. that their larger value for the pion radius might be due to the inevitable model–dependence based on the Born term approach to extract the pion radius. It was also stated that there might be an additional correction obtainable form chiral perturbation theory as it is the case for the nucleon axial radius. Such a correction based on pion loop diagrams had been predicted in 1992 for the axial radius and was verified by the MAMI experiment to a good precision. A list of references pertaining to previous determinations of the nucleon axial radius from (anti)neutrino–proton scattering and pion electroproduction can be found in Ref.. Here, we show that there is indeed a similar kind of correction for the pion radius. This new term modifies the momentum dependence of the longitudinal S–wave amplitude $`L_{0+}^{()}`$ and leads one to expect an even larger pion charge radius than the one given in Eq.(3). It is conceivable that higher order corrections yet to be calculated or contributions from higher multipoles will completely resolve the discrepancy between the pion radius determined from $`\pi e`$ scattering on one side and from charged pion electroproduction on the other. Chiral perturbation theory allows one to make model–independent statements based solely on the spontaneously and explicitly broken chiral symmetry of QCD. S–matrix elements and transition currents are systematically expanded in external momenta and quark (meson) mass insertions, collectively denoted by a small parameter $`q`$. Based on the underlying power counting, to a given order, one has to consider tree as well as pion loop graphs. Of particular interest are the so–called low–energy theorems, which give predictions to a certain order expressed entirely in terms of measurable quantities (a detailed discussion about this issue can be found in Ref.). Our starting point is the low–energy theorem for the longitudinal dipole amplitude $`L_{0+}^{()}`$, accessible in charged pion electroproduction.We use standard notation for the multipoles, the superscript refers to the isospin, the subscripts are $`l\pm `$, with $`l`$ the orbital angular momentum of the pion–nucleon system and the total angular momentum is $`j=l\pm 1/2`$. For a detailed discussion of the kinematics and multipole decomposition in pion electroproduction, see refs.. It has been derived from baryon chiral perturbation theory to third order in the chiral expansion and reads: $`L_{0+}^{()}(\mu ,\nu )`$ $`=`$ $`E_{0+}^{()}(\mu ,\nu )`$ (4) $`+`$ $`{\displaystyle \frac{eg_{\pi N}}{8\pi m}}(\mu ^2\nu )\{{\displaystyle \frac{\kappa _v}{4}}+{\displaystyle \frac{m^2}{6}}r_A^2`$ (5) $`+`$ $`{\displaystyle \frac{\sqrt{(2+\mu )^2\nu }}{2(1+\mu )^{3/2}(\nu 2\mu ^2\mu ^3)}}`$ (6) $`+`$ $`\left({\displaystyle \frac{1}{\nu 2\mu ^2}}{\displaystyle \frac{1}{\nu }}\right)(F_\pi ^V(m^2\nu )1)`$ (7) $`+`$ $`{\displaystyle \frac{m^2}{8\pi ^2F_\pi ^2}}\mathrm{\Xi }_4(\nu \mu ^2)\}+𝒪(q^3),`$ (8) $`E_{0+}^{()}(\mu ,\nu )`$ $`=`$ $`{\displaystyle \frac{eg_{\pi N}}{8\pi m}}\{1\mu +C\mu ^2`$ (9) $`+`$ $`\nu \left({\displaystyle \frac{\kappa _v}{4}}+{\displaystyle \frac{1}{8}}+{\displaystyle \frac{m^2}{6}}r_A^2\right)`$ (10) $`+`$ $`{\displaystyle \frac{\mu ^2m^2}{8\pi ^2F_\pi ^2}}\mathrm{\Xi }_3(\nu \mu ^2)\}+𝒪(q^3),`$ (11) with $`\mathrm{\Xi }_3(\rho )`$ $`=`$ $`\sqrt{1+{\displaystyle \frac{4}{\rho }}}\mathrm{ln}\left(\sqrt{1+{\displaystyle \frac{\rho }{4}}}+{\displaystyle \frac{\sqrt{\rho }}{2}}\right)`$ (12) $`+`$ $`2{\displaystyle _0^1}\sqrt{(1x)[1+x(1+\rho )]}`$ (13) $`\times `$ $`\mathrm{arctan}{\displaystyle \frac{x}{\sqrt{(1x)[1+x(1+\rho )]}}},`$ (14) $`\mathrm{\Xi }_4(\rho )`$ $`=`$ $`{\displaystyle _0^1}𝑑x{\displaystyle \frac{x(12x)}{\sqrt{(1x)[1+x(1+\rho )]}}}`$ (15) $`\times `$ $`\mathrm{arctan}{\displaystyle \frac{x}{\sqrt{(1x)[1+x(1+\rho )]}}},`$ (16) in terms of the dimensionless quantities $`\mu =M_\pi /m`$ and $`\nu =k^2/m^2`$, with $`M_\pi `$ $`(m)`$ the charged pion (nucleon) mass and $`k^20`$ the photon virtuality. Furthermore, $`\kappa _v=\kappa _p\kappa _n=3.71`$ is the isovector anomalous moment of the nucleon, $`F_\pi =92.4`$ MeV the weak pion decay constant, $`g_{\pi N}=13.4`$ the pion–nucleon coupling constant and $`C0.4`$ depends on some low–energy constants. Its precise form is not of interest here but can be found in Ref.. We remark that these expressions have been obtained in a relativistic version of baryon chiral perturbation theory and that some of the kinematical prefactors have not been expanded. However, since we are only after a leading order pion loop effect, the result will not be different from a heavy baryon analysis . For calculating higher order corrections consistently, one either has to use the heavy baryon formalism or the recently proposed infrared regularization . Note also that the dependence of $`L_{0+}^{()}`$ on the axial radius is very weak because the terms $`\nu r_A^2`$ cancel (that is why the axial radius discrepancy discussed before resides in the electric dipole amplitude). As announced, the pion vector form factor $`F_\pi ^V(k^2)`$ appears in the expression for the longitudinal multipole. The pion form factor has the following low–energy expansion, $$F_\pi ^V(k^2)=1+\frac{1}{6}r_\pi ^2_Vk^2+𝒪(k^4).$$ (17) Consequently, to separate the pion radius, one should consider the slope of the longitudinal multipole. For doing that, one has to expand the functions $`\mathrm{\Xi }_{3,4}(\nu \mu ^2)`$ in powers of $`k^2=\nu m^2`$ and pick up all terms proportional to $`\nu `$. This gives: $`{\displaystyle \frac{L_{0+}^{()}}{k^2}}|_{k^2=0}`$ $`=`$ $`{\displaystyle \frac{eg_{\pi N}}{32\pi m}}\{{\displaystyle \frac{1}{M_\pi ^2}}{\displaystyle \frac{1}{mM_\pi }}+{\displaystyle \frac{13}{8m^2}}`$ (18) $`+{\displaystyle \frac{1}{3}}r_\pi ^2_V`$ $`+`$ $`{\displaystyle \frac{1}{32F_\pi ^2}}({\displaystyle \frac{16}{\pi ^2}}1)+𝒪(M_\pi )\}.`$ (19) The first four terms are standard , they comprise the conventional dependence on the pion vector radius, recoil effects and the dominant chiral limit behavior of the slope of the longitudinal multipole. The strong $`1/M_\pi ^2`$ chiral singularity stems from the $`k^2`$–derivative of the pion–pole term which appears already at leading order in the chiral expansion. The last term in Eq.(18) originates from the so–called triangle and tadpole (with three pions coupling to the nucleon at one point) diagrams which are known to play a prominent role in pion photo– and electroproduction (see Fig. 1). The formal reason for the appearance of this new, model–independent contribution at order $`k^2`$ is that one cannot interchange the order of taking the derivative at $`k^2=0`$ and the chiral limit $`M_\pi 0`$. Consequently, all determinations of the pion radius from electroproduction (based on tree-level amplitudes including nucleon and pion form factors) have “measured” the modified radius, $$\stackrel{~}{r}_\pi ^2_V=r_\pi ^2_V+\frac{3}{32F_\pi ^2}\left(\frac{16}{\pi ^2}1\right).$$ (20) The novel term on the right–hand-side of Eq.(20) amounts to 0.266 fm<sup>2</sup>, a bit more than half of the squared pion rms radius, $`r_\pi ^2_V0.44`$fm<sup>2</sup>. Therefore, from the longitudinal multipole alone, one expects to find a larger pion radius if one analyses pion electroproduction based on Born terms, $$\stackrel{~}{r}_\pi ^2_V=(0.44+0.26)\mathrm{fm}^2=(0.83\mathrm{fm})^2,$$ (21) which is even larger than the result of the Mainz analysis, cf. Eq.(3). We point out, however, that the contribution of the pion radius to the derivative of the longitudinal multipole is a factor of ten smaller than the one from the first three terms in the curly brackets in Eq.(18). Therefore, a fourth order analysis is certainly needed to further quantify the “pion radius discrepancy”. Furthermore, the pion form factor contribution to the longitudinal cross section is also present in higher multipoles. In fact, it is known that the convergence of the multipole series for the pion pole term is slow. One should therefore also investigate such effects for these higher multipoles or directly compare the predictions of complete one–loop calculation with the data of the longitudinal electroproduction cross section. For the purpose of demonstrating the significance of chiral loop effects the $`k^2`$-slopes (considered here in Eq.(9) and in Ref. for $`E_{0+}^{()}`$) are, however, best suited. What we have shown here is that as in the case of the nucleon axial mean square radius, the pion loops, which are a unique consequence of the chiral symmetry of QCD, modify the naive Born term analysis and should be taken into account.
warning/0004/hep-th0004126.html
ar5iv
text
# Cosmology in the Randall-Sundrum Brane World Scenario ## Abstract The cosmology of the Randall-Sundrum scenario for a positive tension brane in a 5-D Universe with localized gravity has been studied extensively recently. Here we extend it to more general situations. We consider the time-dependent situation where the two sides of the brane are different AdS/Schwarzschild spaces. We show that the expansion rate in these models during inflation could be larger than in brane worlds with compactified extra dimensions of fixed size. The enhanced expansion rate could lead to the production of density perturbations of substantially larger amplitude. preprint: CLNS 00/1672 Recently Randall and Sundrum presented a static solution to the $`5`$-D (classical) Einstein equations in which spacetime is flat on a 3-brane with positive tension provided that the bulk has an appropriate negative cosmological constant. Even if the fifth dimension is uncompactified, standard $`4`$-D gravity (specifically, Newton’s force law) is reproduced on the brane. In contrast to the compactified case , this follows because the near-brane geometry traps the massless graviton. The extension of their static solution to time-dependent solutions and their cosmological properties have been extensively studied . In this paper, we extend further the time-dependent solution to more general situations. We shall consider only Friedman-Robertson-Walker (FRW) solutions, so the result can be expressed in terms of the Hubble constant. We shall use two approaches to study the problem. One is to solve the Einstein equation straightforwardly and obtain the Hubble constant on the brane. The bulk properties will be encoded into the behavior of the Hubble constant. We shall employ the notation and set-up of Ref . The other approach starts with the solution of the bulk, and the brane is incorporated using the Israel junction condition as first used in this context in Ref . Here we consider the two sides of the brane to be described by two different Anti-deSitter(AdS), or AdS-Schwarzschild (AdSS) spaces, with different cosmological constants and different Newton constants. These cosmological scenarios, including the simplest generalizations of the original Randall-Sundrum brane world that incorporate matter on the brane , all can expand much more rapidly at early times than conventional models based only on 4-D gravity. An important consequence is that the amplitude of primordial density fluctuations produced during inflation could be substantially larger in these scenarios than in brane world models based on compactified extra dimensions of fixed size . We shall consider a $`5`$-dimensional Anti-deSitter spacetime with a positive-tension $`3`$-brane sitting inside. We are interested in the cosmological solutions with a metric of the form $$ds^2=n^2(\tau ,y)d\tau ^2+a^2(\tau ,y)\gamma _{ij}dx^idx^j+b^2(\tau ,y)dy^2.$$ (1) The Einstein equation $`G_{AB}=\kappa ^2T_{AB}=8\pi GT_{AB}`$ can be written as $`G_{00}=3\left[{\displaystyle \frac{\dot{a}}{a}}\left({\displaystyle \frac{\dot{a}}{a}}+{\displaystyle \frac{\dot{b}}{b}}\right){\displaystyle \frac{n^2}{b^2}}\left({\displaystyle \frac{a^{\prime \prime }}{a}}+{\displaystyle \frac{a^{}}{a}}\left({\displaystyle \frac{a^{}}{a}}{\displaystyle \frac{b}{b}}\right)\right)+k{\displaystyle \frac{n^2}{b^2}}\right]`$ (2) $`G_{ij}={\displaystyle \frac{a^2}{b^2}}\gamma _{ij}\left[{\displaystyle \frac{a^{}}{a}}\left({\displaystyle \frac{a^{}}{a}}+2{\displaystyle \frac{n^{}}{n}}\right){\displaystyle \frac{b^{}}{b}}\left({\displaystyle \frac{n^{}}{n}}+2{\displaystyle \frac{a^{}}{a}}\right)+2{\displaystyle \frac{a^{\prime \prime }}{a}}+{\displaystyle \frac{n^{\prime \prime }}{n}}\right]`$ (3) $`+{\displaystyle \frac{a^2}{n^2}}\gamma _{ij}\left[{\displaystyle \frac{\dot{a}}{a}}\left({\displaystyle \frac{\dot{a}}{a}}+2{\displaystyle \frac{\dot{n}}{n}}\right)2{\displaystyle \frac{\ddot{a}}{a}}+{\displaystyle \frac{\dot{b}}{b}}\left(2{\displaystyle \frac{\dot{a}}{a}}+{\displaystyle \frac{\dot{n}}{n}}\right){\displaystyle \frac{\ddot{b}}{b}}\right]k\gamma _{ij}`$ (4) $`G_{05}=3\left({\displaystyle \frac{n^{}}{n}}{\displaystyle \frac{\dot{a}}{a}}+{\displaystyle \frac{a^{}}{a}}{\displaystyle \frac{\dot{b}}{b}}{\displaystyle \frac{\dot{a}^{}}{a}}\right)`$ (5) $`G_{55}=3\left[{\displaystyle \frac{a^{}}{a}}\left({\displaystyle \frac{a^{}}{a}}+{\displaystyle \frac{n^{}}{n}}\right){\displaystyle \frac{b^2}{n^2}}\left({\displaystyle \frac{\dot{a}}{a}}\left({\displaystyle \frac{\dot{a}}{a}}{\displaystyle \frac{\dot{n}}{n}}\right)+{\displaystyle \frac{\ddot{a}}{a}}\right)k{\displaystyle \frac{b^2}{a^2}}\right]`$ (6) where $`\gamma _{ij}`$ is a maximally symmetric $`3`$-dimensional metric, and $`k=1,0,1`$ parametrizes the spatial curvature. The $`3`$-brane is placed at $`y=0`$. On the two sides of the brane $`\left(y<0\right)`$ and $`\left(y>0\right)`$ are two different AdS spaces. The stress-energy-momentum tensor has a bulk and a brane component: $$T_B^A=T_B^A_{bulk}+T_B^A_{brane}.$$ (7) The $`T_B^A_{brane}`$ component corresponds to the matter on the brane placed at $`y=0`$. Since the brane is assumed to be homogeneous and isotropic, this component takes the form $$T_B^A_{brane}=\frac{\delta \left(y\right)}{b}diag(\rho ,p,p,p,0).$$ (8) The matter density on the brane is composed of a cosmological constant term, ordinary matter, radiation, etc. $`T_B^A_{bulk}`$ is the energy-momentum tensor for the bulk matter, which consists of a non-zero cosmological constant (assumed to be different on the different sides of the brane) plus a black hole term, is $$T_B^A_{bulk}=diag(\mathrm{\Lambda }_i,\mathrm{\Lambda }_i,\mathrm{\Lambda }_i,\mathrm{\Lambda }_i,\mathrm{\Lambda }_i)$$ (9) where $`i=+,`$ i.e. $`i=`$ for $`y<0`$ and $`i=+`$ for $`y>0`$. Following Ref, it is convenient to define $`\left[f\right]=f\left(0_+\right)f\left(0_{}\right)`$ to be the jump component for a given function $`f`$ at $`y=0`$, and $`\left\{f\right\}=\left(f^{}\left(0_+\right)+f^{}\left(0_{}\right)\right)/2`$ to be its average component at $`y=0`$. The functions $`n`$, $`a`$, $`b`$, are continuous at the brane, but their derivatives are discontinuous, so the second derivatives will be of the form $$f^{\prime \prime }=f^{\prime \prime }_{\left(y0\right)}+\left[f^{}\right]\delta \left(y\right).$$ (10) Using this notation, the jump part of the Einstein equation for the $`(00)`$, the $`(ij)`$ and the $`(55)`$ components can be written as (the subscript $`0`$ indicates that the functions are evaluated at $`y=0`$) $`{\displaystyle \frac{2\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}{\displaystyle \frac{\left[a^{}\right]}{\left(a_0b_0\right)}}{\displaystyle \frac{\left\{b^{}\right\}}{b_0^2}}{\displaystyle \frac{\left\{a^{}\right\}}{\left(a_0b_0\right)}}{\displaystyle \frac{\left[b^{}\right]}{b_0^2}}={\displaystyle \frac{\kappa ^2}{3}}\left(\mathrm{\Lambda }_+\mathrm{\Lambda }_{}\right)`$ (11) $`{\displaystyle \frac{2\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}+2\left({\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}{\displaystyle \frac{\left\{n^{}\right\}}{n_0b_0}}+{\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left[n^{}\right]}{n_0b_0}}\right)\left({\displaystyle \frac{\left[b^{}\right]}{n_0b_0}}{\displaystyle \frac{\left\{n^{}\right\}}{b_0^2}}+{\displaystyle \frac{\left\{b^{}\right\}}{b_0^2}}{\displaystyle \frac{\left[n^{}\right]}{n_0b_0}}\right)`$ (12) $`2\left({\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}{\displaystyle \frac{\left\{b^{}\right\}}{b_0^2}}+{\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left[b^{}\right]}{b_0^2}}\right)=\kappa ^2\left(\mathrm{\Lambda }_+\mathrm{\Lambda }_{}\right)`$ (13) $`{\displaystyle \frac{2\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}+{\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}{\displaystyle \frac{\left\{n^{}\right\}}{n_0b_0}}+{\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left[n^{}\right]}{n_0b_0}}={\displaystyle \frac{\kappa ^2}{3}}\left(\mathrm{\Lambda }_+\mathrm{\Lambda }_{}\right)`$ (14) while the $`\delta `$-function part of the Einstein equation for the $`(00)`$ and the $`(ij)`$ components can be written as $$\frac{\left[a^{}\right]}{a_0b_0}=\frac{\kappa ^2}{3}\rho ,\frac{\left[n^{}\right]}{n_0b_0}=\frac{\kappa ^2}{3}\left(2\rho +3p\right).$$ (15) The remaining (the average) part of the $`(55)`$ component of the Einstein equation is given by $`{\displaystyle \frac{1}{n_0^2}}\left[{\displaystyle \frac{\dot{a}_0}{a_0}}\left({\displaystyle \frac{\dot{a}_0}{a_0}}{\displaystyle \frac{\dot{n}_0}{n_0}}\right)+{\displaystyle \frac{\ddot{a}_0}{a_0}}\right]={\displaystyle \frac{1}{4}}\left({\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}\right)^2+\left({\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}\right)^2`$ (16) $`+{\displaystyle \frac{1}{4}}{\displaystyle \frac{\left[a^{}\right]}{a_0b_0}}{\displaystyle \frac{\left[n^{}\right]}{n_0b_0}}+{\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}{\displaystyle \frac{\left\{n^{}\right\}}{n_0b_0}}{\displaystyle \frac{\kappa ^2}{3}}{\displaystyle \frac{\mathrm{\Lambda }_++\mathrm{\Lambda }_{}}{2}}{\displaystyle \frac{k}{a_0^2}}.`$ (17) Using the above equations and switching to the proper time of the brane $`t`$ defined by $`d\tau =n(\tau ,0)dt`$, the function $`a_0(t)=a(t,0)`$ describes the evolution of our four-dimensional universe, and will be denoted by $`R\left(t\right)`$. Then the average component of $`G_{55}=\kappa ^2T_{55}`$ gives the Hubble constant $`H=\dot{a_0}/a_0=\dot{R}/R`$ of the brane. The steps to follow are the same as in Ref. After going to the proper time of the brane, the LHS of Eq(16) becomes $$\frac{1}{R}\frac{d^2R}{dt^2}+\frac{1}{R^2}\left(\frac{dR}{dt}\right)^2=\frac{\dot{R}}{R}\frac{d\dot{R}}{dR}+\frac{1}{R^2}\left(\frac{dR}{dt}\right)^2=\frac{R}{2}\left(\frac{dH^2}{dR}\right)+2H^2=\frac{1}{2R^3}\frac{d}{dR}\left(H^2R^4\right).$$ (18) The Hubble constant of the brane can be obtained by integrating the equation $`{\displaystyle \frac{1}{2R^3}}{\displaystyle \frac{d}{dR}}\left(H^2R^4\right)={\displaystyle \frac{\kappa ^4}{36}}\rho \left(\rho +3p\right){\displaystyle \frac{\kappa ^2}{3}}{\displaystyle \frac{\mathrm{\Lambda }_++\mathrm{\Lambda }_{}}{2}}{\displaystyle \frac{k}{R^2}}+\left({\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}\right)^2\left[1+{\displaystyle \frac{3p}{\rho }}\right]`$ (19) $`+\left({\displaystyle \frac{\left\{a^{}\right\}}{a_0b_0}}\right){\displaystyle \frac{\left(\mathrm{\Lambda }_+\mathrm{\Lambda }_{}\right)}{\rho }}.`$ (20) The function $`\left\{a^{}\right\}/a_0b_0`$ is determined by the properties of the bulk. That is, this brane function encodes informations coming from the bulk. If the two sides of the brane are identical AdS spaces, it becomes zero, as required by the $`Z_2`$ symmetry. (All functions depend on $`y`$ only through $`\left|y\right|`$.) The other functions $`\left\{b^{}\right\}/b_0^2`$, $`\left\{n^{}\right\}/n_0b_0`$, $`\left[b^{}\right]/n_0b_0`$, can be determined in terms of $`\left\{a^{}\right\}/a_0b_0`$. Since $`\left[a^{}\right]/a_0b_0`$ and $`\left[n^{}\right]/n_0b_0`$ are already determined by Eq(15), then Eq (11), (12) and (14) allow us to find the three other unknown functions. We may consider expanding $`\left\{a^{}\right\}/a_0b_0`$ in a series of even powers of $`1/R`$: $$\frac{\left\{a^{}\right\}}{a_0b_0}=c_0+\frac{c_1}{R^2}+\frac{c_2}{R^4}+\frac{c_3}{R^6}\mathrm{}$$ (21) We shall see some examples where these coefficients have specific physical interpretations. An alternative way to obtain the Hubble constant has been used in Ref. The bulk is made of two pieces of five-dimensional AdSS space-time separated by the brane. The five-dimensional action is $$S=\frac{1}{16\pi G}_Md^5x\sqrt{g}\left(R+\frac{12}{l^2}\right)+\frac{1}{8\pi G}_Md^4x\sqrt{\gamma }K$$ (22) The cosmological constant used here and in the Ref, $`1/l^2`$, is related to the cosmological constant $`\mathrm{\Lambda }`$ used in Ref by $`1/l^2=\kappa ^2\mathrm{\Lambda }/6`$. The brane separating the two AdSS spaces is described by the equation $`r=R\left(t\right)`$. The extrinsic curvature of the brane, $`K^{ab}=^an^b`$ can be written as $`K_{\mu \nu }=n^c_c\gamma _{\mu \nu }/2`$ where $`\gamma _{\mu \nu }`$ is the induced metric on the brane, and $`n^a`$ is the unit normal of the brane. The indices $`a,b`$ cover the 5D space-time, the indices $`\mu ,\nu `$ cover the 4D space-time, and the indices $`i,j`$ cover the space coordinates of the brane. The junction condition at the brane as required by the Einstein equations is $$\frac{K_{ij}^+}{G_+}\frac{K_{ij}^{}}{G_{}}=8\pi \left(T_{ij}\frac{1}{3}T\gamma _{ij}\right)$$ (23) where $`T_{\mu \nu }`$ is the energy-momentum tensor of matter on the brane, and $`T`$ is its trace. Here we generalize Eq(22) to allow the possibility that the Newton’s constants on the two sides are different. This scenario can happen in string theory, when both the cosmological constant and the Newton’s constant of an AdS space are related to the number of D-branes present. So we may visualize the situation where the two AdS bulk spaces separated by the brane may have different Newton’s constants and cosmological constants. We may also entertain the possibility that the cosmological constants on the two sides are different, a situation that can arise when the brane is a thin-wall approximation of a domain wall . Note that the solution of the jump condition yields a solution to the $`5`$-D Einstein equation, in contrast to the case where the brane is treated as a probe . If the bulk metric has the form $$ds^2=f\left(r\right)dt^2+r^2d\mathrm{\Sigma }_k^2+f^1\left(r\right)dr^2$$ (24) then the velocity vector of the brane $`u^\mu `$, which satisfies $`u^\mu u_\mu =1`$ and $`n^\mu u_\mu =0`$ is given by $`u^t=\left(f+\dot{R}^2\right)^{\frac{1}{2}}f^1`$ and $`u^r=\dot{R}`$, where $`\dot{R}`$ being the derivative with respect to the proper time $`\tau `$. Up to a sign, the unit normal to the brane is given by $`n^t=f^1\dot{R}`$ and $`n^r=\left(f+\dot{R}^2\right)^{\frac{1}{2}}`$. The minus sign is due to the fact that the coordinate $`r`$ is decreasing in the direction $`n^\mu `$. With these components, the spatial components of the extrinsic curvature on the two sides of the brane are $$K_{ij}^{}=\frac{\left(f_{}+\dot{R}^2\right)^{\frac{1}{2}}}{R}\gamma _{ij},K_{ij}^+=\frac{\left(f_++\dot{R}^2\right)^{\frac{1}{2}}}{R}\gamma _{ij}$$ (25) where the relative signs of $`K_{ij}^{}`$ and $`K_{ij}^+`$ follow from the definition of the unit normal $`n^\mu `$. For a matter tensor on the brane of the form given in Eq(8), $`T_{ij}\frac{1}{3}T\gamma _{ij}=(\sigma +\rho _m)/3`$, and the equation describing the evolution of the brane (3-brane) becomes $$\frac{\left(f_{}+\dot{R}^2\right)^{\frac{1}{2}}}{G_{}}+\frac{\left(f_++\dot{R}^2\right)^{\frac{1}{2}}}{G_+}=\frac{8\pi \left(\sigma +\rho _m\right)}{3}R.$$ (26) Using the above equation and the notation $`\stackrel{~}{\lambda }=8\pi \left(\sigma +\rho _m\right)/3`$, the Hubble constant is found to be: $`H^2=\left({\displaystyle \frac{\dot{R}}{R}}\right)^2={\displaystyle \frac{\stackrel{~}{\lambda }^2\left(G_+^2+G_{}^2\right)G_+^2G_{}^2}{\left(G_+^2G_{}^2\right)^2}}+{\displaystyle \frac{\left(f_+G_{}^2f_{}G_+^2\right)}{R^2\left(G_+^2G_{}^2\right)}}`$ (27) $`{\displaystyle \frac{2G_+^2G_{}^2}{\left(G_+^2G_{}^2\right)^2}}\left[\stackrel{~}{\lambda }^4G_+^2G_{}^2+{\displaystyle \frac{\stackrel{~}{\lambda }^2}{R^2}}\left(f_+f_{}\right)\left(G_+^2G_{}^2\right)\right]^{\frac{1}{2}}`$ (28) For an explicit example, we can consider the metric given in Ref, $$f_\pm =k+\frac{R^2}{l_\pm ^2}\frac{\mu _\pm }{R^2}.$$ (29) This reduces to the case considered in Ref if we set $`G_+=G_{}=G`$ and $`l_+=l_{}=l`$. The existence of a horizon at $`r_{h_i}^2=\frac{l_i^2}{2}\left(k\pm \sqrt{k^2+4\mu _i/l_i^2}\right)`$ imposes restrictions on the values of $`\mu _i`$ depending on the value of $`k`$. For $`k=+1`$, a positive value of $`r_{h_i}^2`$ imposes $`\mu _i>0`$ , while for $`k=1`$, the condition is $`\mu _i>l_i^2/4`$. We may also consider a more general metric. Motivated by the generalized AdSS solution of Ref , we can consider a metric of the form $$ds^2=\omega _i^2f_idt^2+\omega _i\left(f_i^1dr^2+r^2d\mathrm{\Omega }_{3,k}\right)$$ (30) where $`f=k\frac{\mu }{r^2}+\frac{r^2}{l^2}\omega ^3`$. In Ref, $`\omega ^3=H_1H_2H_3`$ with $`H_I=1+\frac{q_I}{r^2}`$, where $`q_I`$ are charges. For $`\omega ^3=1`$ and $`G_+=G_{}`$, it reduces to the metric of Ref . In general $`\omega `$ is a function of $`r`$. In this case, we have $$u^r=\dot{R},u^t=\left(\dot{R}^2+\frac{f}{\omega }\right)^{1/2}\frac{\omega ^3}{f},n^t=\frac{\omega ^{3/2}}{f}\dot{R},n^r=\left(\frac{f}{\omega }+\dot{R}^2\right)^{1/2}.$$ (31) It is now straightforward to use the jump condition to obtain an expression for the Hubble constant. To match the two approaches we have discussed, let us go back to the simple case where $`\omega =1`$ and $`G_+=G_{}`$. The equality $`G_+=G_{}=G`$ allows us to define $`\lambda =G\stackrel{~}{\lambda }`$. We can solve for $`\dot{R}`$ and obtain the Hubble constant: $$H^2=\left(\frac{\dot{R}}{R}\right)^2=\frac{\lambda ^2}{4}\frac{f_{}+f_+}{2R^2}+\frac{\left(f_{}f_+\right)^2}{4R^4\lambda ^2}$$ (32) Using the expressions for the functions $`f_i`$, the Hubble constant becomes $`H^2={\displaystyle \frac{\lambda ^2}{4}}{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{1}{l_{}^2}}+{\displaystyle \frac{1}{l_+^2}}\right)+{\displaystyle \frac{1}{4\lambda ^2}}\left({\displaystyle \frac{1}{l_{}^2}}{\displaystyle \frac{1}{l_+^2}}\right)^2+{\displaystyle \frac{k}{R^2}}`$ (33) $`+{\displaystyle \frac{1}{R^4}}\left\{{\displaystyle \frac{\mu _{}+\mu _+}{2}}{\displaystyle \frac{\mu _+\mu _{}}{2\lambda ^2}}\left({\displaystyle \frac{1}{l_{}^2}}{\displaystyle \frac{1}{l_+^2}}\right)\right\}+{\displaystyle \frac{1}{R^8}}{\displaystyle \frac{\left(\mu _+\mu _{}\right)^2}{4\lambda ^2}}.`$ (34) To compare Eq(19) with this expression, we must first integrate Eq(19), which will generate an integration constant. We see that the integration constant is $$C=\frac{\mu _{}+\mu _+}{2}\frac{\mu _+\mu _{}}{2\lambda ^2}\left(\frac{1}{l_{}^2}\frac{1}{l_+^2}\right)$$ (35) where we identify $$\lambda =\frac{\kappa ^2\sigma }{3},\frac{1}{l_i^2}=\frac{\kappa ^2\mathrm{\Lambda }_i}{6}$$ (36) and $`c_0=\kappa ^2\left(\mathrm{\Lambda }_+\mathrm{\Lambda }_{}\right)/12,c_1=0,c_2=\left(\mu _+\mu _{}\right)/2\lambda `$. The integration constant was obtained in Ref, and the $`R^8`$ term was first obtained in Ref. The $`\mu `$ term has been interpreted as due to a $`N=4`$ super-Yang-Mills theory on the brane via the AdS/CFT correspondence . The holographic principle states that information in the bulk is encoded in the data on the boundary. Intuitively, one may understand this as a boundary value problem. Evolution of the boundary to the bulk fixes the properties of the bulk in terms of the values at the boundary. However, this intuition does not apply to the brane world scenario when there are different AdSS spaces on the two sides of the brane. It is natural to ask if data on the brane encodes all properties of both bulks, or only a combination of the bulk information with no chance to dis-entangle them. The presence of the $`(\mu _+\mu _{})^2/R^8`$ term in addition to the $`(\mu _++\mu _{})^2/R^4`$ term in the Hubble constant equation allows one to determine both $`\mu _+`$ and $`\mu _{}`$. That is, there are two different conformal field theories on the brane, which couple to each other. This suggests that there is enough information on the brane to determine fully all the bulk properties on each side of the brane. In this sense, the holographic principle is deep. Tuning the effective cosmological constant in the brane to zero requires $$\frac{\lambda ^2}{4}\frac{1}{2}\left(\frac{1}{l_{}^2}+\frac{1}{l_+^2}\right)+\frac{1}{4\lambda ^2}\left(\frac{1}{l_{}^2}\frac{1}{l_+^2}\right)^2=0$$ (37) or using $`\lambda =\frac{\kappa ^2\sigma }{3}`$, $$\frac{\kappa ^4\sigma ^2}{36}\frac{\kappa ^2\left(\mathrm{\Lambda }_{}+\mathrm{\Lambda }_+\right)}{12}+\frac{\left(\mathrm{\Lambda }_{}\mathrm{\Lambda }_+\right)^2}{16\sigma ^2}=0.$$ (38) For $`\mathrm{\Lambda }_{}=\mathrm{\Lambda }_+`$, this reduces to the Randall-Sundrum case. Otherwise, the equation has the two solutions $$\kappa ^2\sigma ^2=\frac{3}{2}\left(\sqrt{\mathrm{\Lambda }_{}}\pm \sqrt{\mathrm{\Lambda }_+}\right)^2.$$ (39) the $`{}_{}{}^{\prime \prime }+_{}^{\prime \prime }`$ solution being the one found in Ref. Note that, if we choose either $`\mathrm{\Lambda }`$ to be zero, the two solutions for the value of the brane tension merge, and the minimum of $`H^2`$ is exactly zero. By choosing $`\mu _i=0`$ and $`\rho _m=0`$ in Eq(33) we obtain $$H^2=\frac{\kappa ^4\sigma ^2}{36}\frac{\kappa ^2\mathrm{\Lambda }}{12}+\frac{\mathrm{\Lambda }^2}{16\sigma ^2}=\left(\frac{\kappa ^2\sigma }{6}\frac{\mathrm{\Lambda }}{4\sigma }\right)^2$$ (40) This intriguing property implies that the minimum of the $`4`$-dimensional effective cosmological constant is bounded below by zero. The relationship between the 5D and 4D Newton constants is $$\frac{l_+}{G_+}+\frac{l_{}}{G_{}}=\frac{2}{G_{\left(4\right)}}.$$ (41) This can be obtained by dimensional reduction. We may also identify the 4D Newton constant by adding matter to the brane, expanding for $`\rho _m\sigma `$ and identifying the coefficient of $`\rho _m`$ in Eq(33). This shows that conventional cosmology can be recovered at large enough values of $`R`$, where the matter density is low and the additional terms in Eq(33) become small. Choosing the “$``$” solution of Eq(39) does not, however, lead to conventional cosmology at large enough $`R`$. It is instructive to consider a simplified version of Eq(33) to better appreciate its significance. To be concrete, let us assume that $`k=0=\mu _+\mu _{}`$, but still allow $`l_{}l_+`$. Then we find that $$H^2=\frac{(\lambda ^2\lambda _+^2)(\lambda ^2\lambda _{}^2)}{4\lambda ^2}+\frac{\mu }{R^4},$$ (42) where $`\mu =\mu _\pm `$ and $`\lambda _\pm ^2=(1/l_+\pm 1/l_{})^2`$; quite generally, this is to be augmented by an equation governing the evolution of $`\lambda `$, which can always be written schematically as $`\dot{\lambda }=3H(1+w)\lambda `$, where for mixtures of matter, radiation, and evolving fields, $`1w1`$, so $`\lambda `$ decreases with expansion generically. It is easy to see that the first term in Eq(42) is negative for any $`\lambda ^2<\lambda _+^2`$, which means that for such values of $`\lambda `$, there are only sensible cosmological models if $`\mu >0`$. In such a case, one expects that the Universe expands to a maximum $`R`$ and then recollapses, generically, although it is conceivable that the parameters could be fine-tuned to avoid recollapse and attain $`\lambda \lambda _{}`$ as $`R\mathrm{}`$. However, it is clear that the $`\mu /R^4`$ term plays an essential role in the expansion rate throughout the history of such models, whether they collapse or expand forever, and consequently conventional cosmology is not recovered in any limit. Thus, such models could never reproduce the successes of cosmological nucleosynthesis theory, for example, and would not yield acceptable theories of large scale structure, even if expansion to very large (or infinite) $`R`$ is possible. On the other hand, for $`\lambda >\lambda _+`$, the Universe always expands, irrespective of the sign of $`\mu `$, although, for $`\mu <0`$ it expands from a minimum value of $`R`$. As long as there is some component in $`\lambda `$ with $`w<1/3`$, such as the brane tension or the vacuum energy density of a brane field, the $`\mu /R^4`$ tern in Eq(42) becomes progressively less important as the Universe expands, and conventional cosmology can be recovered provided that $`\lambda \lambda _+`$ as $`R\mathrm{}`$. During the inflation era, the matter density, $`\rho _m`$, is dominated by the inflaton, which we will take to be a single component scalar field $`\varphi `$ with some effective potential $`V(\varphi )`$. The inflaton will tend to roll toward its potential minimum, and, while the field is rolling slowly, the energy density in the field is dominated by $`V(\varphi )`$, which decreases with time as the Universe expands, but only slowly if $`V(\varphi )`$ is flat enough. If the inflation commences with $`\lambda >1/l_++1/l_{}`$, then $`H^2`$ decreases as $`\varphi `$ rolls toward the minimum of $`V(\varphi )`$, where $`V(\varphi )=V_{min}`$. Provided that $`V_{min}+\sigma `$ is tuned so that $`\kappa ^2(V_{min}+\sigma )/3=\lambda _+`$, conventional cosmology can be recovered. <sup>§</sup><sup>§</sup>§Note that to the extent that we may think of $`\sigma `$ as the tension of a domain wall solution for self-gravitating supergravity fields off the brane, and $`V_{min}`$ as arising from standard model fields that are confined to the brane, the precise cancellation of the cosmological constant involves cooperation between brane and bulk physical fields. One of the intriguing features of Eq(33) is that for $`\lambda `$ well above $`1/l_+^2+1/l_{}^2`$, the expansion rate grows linearly with $`\lambda `$, i.e. $`H\lambda /2`$ (assuming that the other terms in Eq(33) that are proportional to inverse powers of $`R^2`$ can be neglected).This is also true in scenarios where the AdS spaces on either side of the brane are identical. With $`\rho _mV(\varphi )`$, this limit applies when $`V(\varphi )`$ is at least a factor of a few larger than $`\sigma _+=\sqrt{3/2}(\sqrt{\mathrm{\Lambda }_+}+\sqrt{\mathrm{\Lambda }_{}})/\kappa `$. The expansion rate during this epoch, $`H\kappa ^2V(\varphi )/6`$, can be much larger than the expansion rate in conventional inflation for the same value of $`V(\varphi )`$, which is $`H_{conv}=(8\pi V(\varphi )/3M_P^2)^{1/2}`$, where $`M_P`$ is the Planck mass: $`H/H_{conv}\kappa ^2M_P\sqrt{V(\varphi )/96\pi }=\sqrt{V(\varphi )/8\sigma _+}(\sqrt{l_+/l_{}}+\sqrt{l_{}/l_+})`$. Conceivably, $`HH_{conv}`$ only as $`\varphi `$ settles into its potential minimum, and during much of inflation $`H\kappa ^2V(\varphi )/6`$ instead. This can happen in a number of ways. First, if $`V(\varphi )`$ has some characteristic scale $`V_0`$, which can occur if the effective potential is flat until it plummets to its minimum at $`V_{min}V_0`$ (or $`V_{min}=0`$), and $`V_0\sigma _+`$, then $`H\kappa ^2V_0/6`$ for most of the inflationary era. Second, if $`V(\varphi )`$ is an increasing, polynomial function of $`\varphi `$, for example a simple powerlaw $`V(\varphi )\varphi ^n`$, then $`H=\kappa ^2V(\varphi )/6`$ at sufficiently large values of $`\varphi `$, and the conventional expansion rate is only achieved when $`V(\varphi )\sigma _+`$. Finally, when either $`l_+`$ or $`l_{}`$ is much larger than the other, then the expansion rate can be $`H\kappa ^2V(\varphi )/6`$ even when $`V(\varphi )`$ is only a factor of a few larger than $`\sigma _+`$. Generally speaking, this enhanced expansion rate would have no consequence for the mean properties of the Universe; reheating and the end of inflation would be unaffected. But, the production of density perturbations from fluctuations in the inflaton field could be dramatically different than in models for brane world inflation with compactified extra dimensions of fixed size, where $`H=H_{conv}`$ and $`\delta \rho /\rho `$ generally is too small. To evaluate the differences, suppose that $`\sigma =\sigma _+`$ and $`V_{min}=0`$. We estimate the magnitude of the density fluctuations produced from the usual relationship $`(\delta \rho /\rho )_q(V^{}(\varphi )H/\dot{\varphi }^2)_{h.c.}(H^3/V^{})_{h.c.}`$, where $`V^{}=V(\varphi )/\varphi `$, and the subscript “h.c.” means that we evaluate $`H^3/V^{}`$ when a perturbation of comoving scale $`q^1`$ crosses the horizon during inflation . Two simple examples will suffice to illustrate the effect. If $`V(\varphi )=m^2\varphi ^2/2`$, an example of “chaotic inflation” (e.g. ), then we estimate $`\delta \rho /\rho \kappa m^{3/2}N_q^{5/4}`$, which is to be compared to the value $`(\delta \rho /\rho )_{conv}(m/M_P)N_q`$ for $`H=H_{conv}`$, where $`N_q`$ is the number of e-foldings of $`R(t)`$ remaining in inflation after horizon crossing for a comoving scale $`q^1`$. The fluctuations on scales that cross the horizon while $`H\kappa ^2V(\varphi )/6`$ differ from what one would get for $`H=H_{conv}`$ by a factor $`\kappa m^{1/2}M_PN_q^{1/4}`$. If instead $`V(\varphi )=V_0[1\mathrm{exp}(\varphi /m)]`$, which arises in “brane inflation” (e.g. , ), then we estimate $`\delta \rho /\rho \kappa ^2V_0N_q/m`$, which differs from the estimate $`(\delta \rho /\rho )_{conv}V_0^{1/2}N_q/M_Pm`$ found for $`H=H_{conv}`$ by a factor $`\kappa ^2V_0^{1/2}M_P`$. If the characteristic energy scales of these inflationary models are comparable to $`\kappa ^{2/3}`$, and $`M_P\kappa ^{2/3}`$, then the implied density fluctuations are much larger than for $`H=H_{conv}`$ with the same inflaton potentials. The enhancement in $`\delta \rho /\rho `$ for these two models is a consequence of the faster expansion rate during inflation, so that the $`H^3`$ factor in our estimate of $`\delta \rho /\rho `$ is larger than $`H_{conv}^3`$, but is mitigated by the evolution of $`\varphi `$, which tends to raise the value of $`V^{}`$ at horizon crossing for these potentials. There are also potentials for which there is no effect, and for which the suppression due to $`V^{}`$ outweighs the enhancement due to $`H^3`$. Thus, for $`V(\varphi )=\lambda \varphi ^4/4`$ or $`V(\varphi )=V_0\lambda \varphi ^4/4`$. we estimate $`\delta \rho /\rho \lambda ^{1/2}N_q^{3/2}`$, just as in conventional inflation, and for $`V(\varphi )=m^{4n}\varphi ^n/n`$ with $`n>4`$ the density fluctuations may even be smaller than for $`H=H_{conv}`$. A similar effect is seen in brane world cosmologies in which the extra dimensions are compactified, provided that the extra dimensions are smaller, and therefore the effective Newton “constant” is larger, during inflation than today . We thank Eanna Flanagan and Vatche Sahakian for useful discussions. The research of S.-H.H.T. is partially supported by the National Science Foundation. I.W. gratefully acknowledges support from NASA.
warning/0004/cond-mat0004241.html
ar5iv
text
# References ## References
warning/0004/cond-mat0004225.html
ar5iv
text
# Tricritical behavior in deterministic aperiodic Ising systems ## Abstract We use a mixed-spin model, with aperiodic ferromagnetic exchange interactions and crystalline fields, to investigate the effects of deterministic geometric fluctuations on first-order transitions and tricritical phenomena. The interactions and the crystal field parameters are distributed according to some two-letter substitution rules. From a Migdal-Kadanoff real-space renormalization-group calculation, which turns out to be exact on a suitable hierarchical lattice, we show that the effects of aperiodicity are qualitatively similar for tricritical and simple critical behaviour. In particular, the fixed point associated with tricritical behaviour becomes fully unstable beyond a certain threshold dimension (which depends on the aperiodicity), and is replaced by a two-cycle that controls a weakened and temperature-depressed tricritical singularity. The introduction of quenched disorder weakens (and sometimes eliminates) first-order transitions and tricritical singularities in the phase diagram of statistical models. In two dimensions, rigorous arguments show that any amount of disorder completely eliminates first-order transitions in ferromagnetic model systems . In three dimensions, approximate real-space and perturbative renormalization-group analyses , as well as numerical simulations, indicate that a finite strength of disorder is required to weaken first-order transitions and depress the tricritical temperature. In particular, disordered versions of the two-dimensional ferromagnetic $`q`$-state Potts model (for $`q>4`$, on the square lattice, the uniform model displays a first-order transition) and the Blume-Emery-Griffiths (BEG) model (whose uniform version displays tricritical and critical-end points), which have been thoroughly investigated, are well adjusted to this scenario. It remains unclear the important question of what are (if any) the universality classes of the disorder-induced continuous transitions in these systems (see, for example, a recent review by Cardy ). Instead of looking at the (presumably) more difficult problem posed by fluctuations associated with quenched disorder, in the present publication we consider the effects on first-order transitions of the geometric fluctuations introduced by deterministic but aperiodically distributed exchange interactions. For quenched disordered interactions, the Harris criterion indicates a change in the critical behaviour of simple ferromagnets whenever the critical exponent associated with the singularity of the specific heat of the underlying uniform model is positive. According to a heuristic argument of Luck , which has been checked in a number of cases \[12, and references therein\], the introduction of aperiodic interactions leads to an analogous criterion of relevance of the geometric fluctuations on the critical (second-order) behaviour. Some of us have recently shown that a similar criterion may be exactly established for ferromagnetic Potts models on Migdal-Kadanoff hierarchical lattices, with a layered distribution of exchange interactions according to a class of two-letter substitution rules . For relevant geometric fluctuations, there appears a two-cycle of the recursion relations in parameter space that gives rise to a new universality class of (aperiodic) critical behaviour . Along the lines of these investigations, we introduce aperiodic interactions in a simple mixed-spin model to analyze the effects on tricritical behaviour and first-order phase boundaries. It should be mentioned that recent extensive Monte Carlo calculations indicate that the phase transition of the $`8`$-state square-lattice Potts model is indeed driven to second order by a layered aperiodic distribution of exchange couplings . Besides the better known BEG model, another simple generalization of the Ising model displaying first-order transitions and tricritical points is a mixed-spin system, given by the Hamiltonian $$=\underset{(i,j)}{}J_{ij}\sigma _iS_j+\underset{j}{}D_jS_j^2,$$ (1) where $`\sigma _i=\pm 1`$, for $`i`$ belonging to one sublattice of a bipartite lattice, $`S_j=\pm 1`$ or $`0`$, for $`j`$ belonging to the other sublattice, and the first sum is over nearest-neighbor sites on different sublattices. The description of this mixed-spin model demands a larger unit cell than the BEG model in zero field. This extra difficulty, however, poses no problem to the study of the model under a Migdal-Kadanoff real-space renormalization-group approximation. Moreover, we need just two scaling fields, instead of three as in the BEG model, to describe the even space of parameters. We then take advantage of this model, and of the simplicity of the Migdal-Kadanoff approximation (which turns out to be exact on a suitable hierarchical lattice), to introduce aperiodic interactions for analyzing the effects of geometric fluctuations on the main features of the phase diagram. To obtain the Migdal-Kadanoff recursion relations, it is convenient to rewrite Hamiltonian (1) in the equivalent form $$=\underset{(i,j)}{}J_{ij}\sigma _iS_j+\frac{1}{2}\underset{(i,j)}{}D_{ij}\sigma _i^2S_j^2.$$ (2) Many (mainly approximate) results are known for the uniform ($`J_{ij}=J`$, $`D_{ij}=D`$) version of this model. On a honeycomb lattice, a star-triangle transformation (summing over $`S_j`$) can be used to reduce the problem to an exactly soluble spin-$`1/2`$ Ising model on a simple triangular lattice, in which case, however, the temperature $`(k_BT)`$ versus “anisotropy” $`(D/J)`$ phase diagram presents only a line of continuous transitions . For the so-called union jack lattice, an exact solution can also be found for a restricted range of parameters, by mapping the model onto an eight-vertex problem . On a lattice of sufficiently high coordination, some effective-field and self-consistent approximations, as well as a Bethe lattice calculation , suggest the existence of a first-order boundary that becomes a $`\lambda `$ line beyond a tricritical point. A detailed Migdal-Kadanoff renormalization-group calculation predicts the existence of a tricritical point on hypercubic lattices of dimension $`d2.1`$, which precludes the case of planar lattices. This is also confirmed by renormalization-group calculations in momentum space, at one-loop approximation, that do support the existence of the tricritical point predicted by the Curie-Weiss version of the model. Monte Carlo and numerical transfer matrix calculations point out in this direction as well, with no indication of tricritical phenomena in this mixed-spin model on a square lattice. We now consider Hamiltonian (2) and suppose that $`J_{ij}`$ (and $`D_{ij}`$) may assume one out of two values, $`J_A`$ or $`J_B`$ ($`D_A`$ or $`D_B`$), according to the sequence of letters generated by the iteration of a substitution rule. In this paper, we work with two distinct binary rules, $$\left(i\right)AABB,BAAA,$$ (3) and $$\left(ii\right)AAAB,BAAA.$$ (4) For example, the iteration of the first rule leads to the following stages, $$AABBABBAAAAAA\mathrm{}.$$ (5) Each rule is characterized by a substitution matrix, which relates the number of letters $`A`$ and $`B`$ in one stage of the iteration with the corresponding numbers in the previous stage. For the first rule, the substitution matrix is given by $`𝐌=\left(\begin{array}{cc}1\hfill & 3\hfill \\ 2\hfill & 0\hfill \end{array}\right),`$ with eigenvalues $`\lambda _1=3`$ and $`\lambda _2=2`$. For the second rule, we have $`𝐌=\left(\begin{array}{cc}2\hfill & 3\hfill \\ 1\hfill & 0\hfill \end{array}\right),`$ with eigenvalues $`\lambda _1=3`$ and $`\lambda _2=1`$. In a given stage of the iteration, the fluctuation in the number of letters $`A`$ or $`B`$ relative to the mean number behaves asymptotically as $`N^\omega `$, where $`N`$ is the total number of letters in the sequence, and $`\omega =\mathrm{ln}\left|\lambda _2\right|/\mathrm{ln}\lambda _1`$ is a wandering exponent. We thus see that the first rule, with an wandering exponent $`\omega =\mathrm{ln}2/\mathrm{ln}3`$, gives rise to stronger geometric fluctuations than the second one, with $`\omega =0`$, and should be more effective in perturbing the critical behaviour. The Migdal-Kadanoff (MK) approximation on a $`d`$-dimensional hypercubic lattice turns out to be exact on the hierarchical cell shown in figure 1. For this type of cell, the MK scheme corresponds to a decimation of the two spins located along each bond, a spin-$`1/2`$ and a spin-$`1`$, followed by the moving and collapsing of $`m`$ such bonds. There is also a relationship, $`d=1+\mathrm{ln}m/\mathrm{ln}3`$, between the number of branches $`m`$ and the Euclidean dimension $`d`$. Note that the renormalization procedure amounts to a reverse application of the substitution rule generated by the aperiodic sequence. Also, note that we can as well perform the more usual trick of bond-moving before decimation (which is also exact on a suitable hierarchical lattice), but the qualitative results should not depend on this choice. For the first aperiodic rule, $`AABB`$, $`BAAA`$, the MK procedure yields the recursion relations $`K_A^{}=`$ $`{\displaystyle \frac{m}{2}}\mathrm{ln}\left\{\mathrm{exp}(K_A)+\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_B\right)+\mathrm{cosh}\left(K_B\right)\mathrm{exp}\left({\displaystyle \frac{\mathrm{\Delta }_A}{2}}+{\displaystyle \frac{\mathrm{\Delta }_B}{2}}\right)\right\}`$ (7) $`+{\displaystyle \frac{m}{2}}\mathrm{ln}\left\{\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_B\right)+\mathrm{exp}(K_A)+\mathrm{cosh}\left(K_B\right)\mathrm{exp}\left({\displaystyle \frac{\mathrm{\Delta }_A}{2}}+{\displaystyle \frac{\mathrm{\Delta }_B}{2}}\right)\right\},`$ $`K_B^{}=`$ $`{\displaystyle \frac{m}{2}}\mathrm{ln}\left\{\mathrm{exp}(K_A)+\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_A\right)+\mathrm{cosh}\left(K_A\right)\mathrm{exp}\left(\mathrm{\Delta }_A\right)\right\}`$ (9) $`+{\displaystyle \frac{m}{2}}\mathrm{ln}\left\{\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_A\right)+\mathrm{exp}(K_A)+\mathrm{cosh}\left(K_A\right)\mathrm{exp}\left(\mathrm{\Delta }_A\right)\right\},`$ $`\mathrm{\Delta }_A^{}=`$ $`m\mathrm{ln}\{\mathrm{exp}(\mathrm{\Delta }_B)`$ (13) $`\times \left[\mathrm{exp}(K_A)+\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_B\right)+\mathrm{cosh}\left(K_B\right)\mathrm{exp}\left({\displaystyle \frac{\mathrm{\Delta }_A}{2}}+{\displaystyle \frac{\mathrm{\Delta }_B}{2}}\right)\right]`$ $`\times \left[\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_B\right)+\mathrm{exp}(K_A)+\mathrm{cosh}\left(K_B\right)\mathrm{exp}\left({\displaystyle \frac{\mathrm{\Delta }_A}{2}}+{\displaystyle \frac{\mathrm{\Delta }_B}{2}}\right)\right]`$ $`\times [2\mathrm{cosh}(K_A)\mathrm{cosh}(K_B)+\mathrm{exp}({\displaystyle \frac{\mathrm{\Delta }_A}{2}}+{\displaystyle \frac{\mathrm{\Delta }_B}{2}})]^2\},`$ $`\mathrm{\Delta }_B^{}=`$ $`m\mathrm{ln}\{\mathrm{exp}(\mathrm{\Delta }_A)`$ (17) $`\times \left[\mathrm{exp}(K_A)+\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_A\right)+\mathrm{cosh}\left(K_A\right)\mathrm{exp}\left(\mathrm{\Delta }_A\right)\right]`$ $`\times \left[\mathrm{exp}(K_A)\mathrm{cosh}\left(2K_A\right)+\mathrm{exp}(K_A)+\mathrm{cosh}\left(K_A\right)\mathrm{exp}\left(\mathrm{\Delta }_A\right)\right]`$ $`\times [2\mathrm{cosh}^2(K_A)+\mathrm{exp}\left(\mathrm{\Delta }_A\right)]^2\},`$ where $`K_{A,B}=\beta J_{A,B}`$ and $`\mathrm{\Delta }_{A,B}=\beta D_{A,B}`$, with $`\beta =1/k_BT`$. For all values of $`m1`$, these recursion relations have a set of trivial fixed points, given by $$P_+(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})=(0,0,\mathrm{},\mathrm{}),$$ (18) which is a sink of high-density paramagnetic phase (where the density is related to the mean value $`S_i^2`$, so that low density means the predominance of spin $`0`$), $$P_{}(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})=(0,0,\mathrm{},\mathrm{}),$$ (19) which is a sink of low-density paramagnetic phase, $$O(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})=(0,0,0,0),$$ (20) corresponding to the high-temperature boundary between the paramagnetic phases, $$F_+(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})=(\mathrm{},\mathrm{},\mathrm{},\mathrm{}),$$ (21) associated with a zero-temperature high-density ferromagnetic phase, and $$F_{}(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})=(\mathrm{},\mathrm{},\mathrm{},\mathrm{}),$$ (22) with $`\mathrm{\Delta }_{A,B}^{}/K_{A,B}^{}=2`$, which corresponds to a zero-temperature low-density ferromagnetic phase. For $`m>1`$ (which corresponds to $`d>1`$), there is a non-trivial fixed point at $`\mathrm{\Delta }_A^{}=\mathrm{\Delta }_B^{}=\mathrm{}`$, for $`K_A^{}=K_B^{}`$ finite. This fixed point, which we shall call $`I`$, is associated with the critical behaviour of the simple spin-$`1/2`$ Ising model, since fixing the crystal field (biquadratic exchange) at $`\mathrm{}`$ completely prevents the $`S`$ spins to assume $`0`$ values. In this spin-$`1/2`$ space, it should be pointed out that the recursion relations also present a two-cycle (that is, a set of two fixed points of the second-iterate). As discussed in a recent publication , this two-cycle is associated with the new universality class of the aperiodic ferromagnetic Ising model. Indeed, the critical behaviour of the Ising model on the hierarchical lattice underlying the MK approximation, and with aperiodic interactions according to the rule $`AABB`$, $`BAAA`$, is controlled by this two-cycle, the singularity being weaker as compared with the uniform (periodic) model. For $`m3.33`$ (numerical evidence points in fact to $`10/3`$), corresponding to $`d2.1`$ (or $`\mathrm{ln}10/\mathrm{ln}3=2.0959\mathrm{}`$, as suggested by the numerical calculations), there appear two novel non-trivial fixed points. In the uniform case ($`J_A=J_B,`$ $`D_A=D_B`$), Quadros and Salinas have shown that one of them is a discontinuity fixed point, associated with a first-order phase transition (according to an application of the Nienhuis and Nauenberg criterion for the identification of discontinuity fixed points). On the basis of a detailed analysis of the connectivity of the flow lines in parameter space, the other fixed point of the uniform model was shown to be associated with the tricritical behaviour. In the uniform case, the discontinuity fixed point displays a saddle character, with an attractive manifold emerging from the zero-temperature high-density ferromagnetic trivial fixed point, and from the tricritical fixed point (see figure 2, where a schematic view of the flows is presented). The repulsive directions flow towards the sink of the low-density paramagnetic phase, and towards the zero-temperature low-density ferromagnetic trivial fixed point. The stable manifold is thus associated with a first order line. The fully unstable tricritical fixed point is connected with the Ising, $`D=\mathrm{}`$, fixed point, giving rise to a second-order boundary. In the present aperiodic case, for $`m3.33`$, these two fixed points still appear in the $`K_A=K_B`$, $`\mathrm{\Delta }_A=\mathrm{\Delta }_B`$, subspace, but since the parameter space is four-dimensional we have to be careful to generalize the overall picture of the last paragraph. In fact, we should pay attention to the fact that now there are four scaling fields: the reduced temperature, two crystal fields (instead of only one, as in the uniform case), and the strength of the aperiodicity, as measured, for example, by the ratio $`r=J_A/J_B`$. This last scaling field, which is of completely different nature, is the determinant factor for the analysis of stability of the fixed points against aperiodicity. If it turns out to be relevant around a certain fixed point, it means that this fixed point cannot be reached unless $`r`$ assumes a well-defined value, usually unity, so that any amount of aperiodicity changes the critical behaviour controlled by this particular fixed point. On the other hand, irrelevancy of this scaling field suggests that, whatever the strength of the geometric fluctuations, this fixed point continues to control the critical behaviour. Consider now $`m=3.4`$ (dimension $`d2.1`$). The fixed points are at $`K_A^{}=K_B^{}=1.5248\mathrm{}`$, $`\mathrm{\Delta }_A^{}=\mathrm{\Delta }_B^{}=2.2118\mathrm{}`$ (discontinuity fixed point in the uniform model) and $`K_A^{}=K_B^{}=1.1447\mathrm{}`$, $`\mathrm{\Delta }_A^{}=\mathrm{\Delta }_B^{}=1.3953\mathrm{}`$ (tricritical fixed point of the uniform case). The linearization of the recursion relations (7)-(17), in the neighborhood of this fixed point lead to the eigenvalues $`\mathrm{\Lambda }_1=4.3687\mathrm{},\mathrm{\Lambda }_2=0.7705\mathrm{},\mathrm{\Lambda }_3=3.8520\mathrm{},\mathrm{\Lambda }_4=0.5825\mathrm{}.`$ The first two eigenvalues are the same as in the uniform case. The modulus of the last two eigenvalues indicates the preservation of the saddle character of this fixed point. The linearization about the second fixed point yields the eigenvalues $`\mathrm{\Lambda }_1=4.1671\mathrm{},\mathrm{\Lambda }_2=1.2360\mathrm{},\mathrm{\Lambda }_3=3.7346\mathrm{},\mathrm{\Lambda }_4=0.9194\mathrm{}.`$ Again, $`\mathrm{\Lambda }_1`$ and $`\mathrm{\Lambda }_2`$ assume the same values of the uniform case. The modulus of the third eigenvalue is larger than $`1`$, so far preserving the unstable character of the uniform tricritical fixed point. The last eigenvalue, however, indicates an attractive direction towards this fixed point, in contrast to the uniform model, in which case the tricritical fixed point is fully unstable. However, we have already remarked that a given characteristic fixed point associated with the uniform model still controls the critical behaviour of the aperiodic system only if the strength of the aperiodicity is an irrelevant scaling field. This is precisely what is happening in this case. The modulus of $`\mathrm{\Lambda }_4`$ asserts that the tricritical fixed point associated with the uniform model can be reached, even in the presence of aperiodicity. An analysis of the flow lines of the recursion relations in parameter space fully supports the idea that these two fixed points continue to perform exactly the same functions in the aperiodic as well as in the uniform case. In other words, the first one is indeed a discontinuity fixed point, and the other one is associated with the tricritical behaviour. In a phase diagram consisting of temperature, crystal fields and the aperiodicity ratio $`r`$, there exists then a line of tricritical points extending along the $`r`$ direction. Finally, we note that $`\mathrm{\Lambda }_3`$ and $`\mathrm{\Lambda }_4`$ are negative, for both fixed points. This is a common situation for these aperiodic systems , which reflects a flipping approximation to (or furthering from) the fixed points, and is related to the discrete nature of the renormalization procedure and to the existence of two competing energy scales, $`A`$ and $`B`$. For larger dimensions (larger values of $`m`$), this whole picture is changed. Take, for example, $`m=9`$, corresponding to three dimensions. The discontinuity fixed point of the uniform model is located at $`K_A^{}=K_B^{}=4.8427\mathrm{}`$, $`\mathrm{\Delta }_A^{}=\mathrm{\Delta }_B^{}=9.0255\mathrm{}`$. The tricritical fixed point is located at $`K_A^{}=K_B^{}=0.4514\mathrm{}`$, $`\mathrm{\Delta }_A^{}=\mathrm{\Delta }_B^{}=0.2190\mathrm{}`$. The linearization of the recursion relations in the neighborhood of these fixed points lead to the eigenvalues $`\mathrm{\Lambda }_1=12.0154\mathrm{},\mathrm{\Lambda }_2=0.0032\mathrm{},\mathrm{\Lambda }_3=10.5075\mathrm{},\mathrm{\Lambda }_4=0.0002\mathrm{},`$ around the discontinuity fixed point (the first two eingenvalues are the same as in the uniform case), and $`\mathrm{\Lambda }_1=9.1600\mathrm{},\mathrm{\Lambda }_2=2.5865\mathrm{},\mathrm{\Lambda }_3=9.0707\mathrm{},\mathrm{\Lambda }_4=1.7413\mathrm{},`$ around the second fixed point (again, the first two eingenvalues are the same as in the uniform case). Note that this time the modulus of all of the eigenvalues of the uniform tricritical fixed point are larger than unity, which means that it is fully unstable against any aperiodic perturbations. Hence, it cannot be reached whatever the strength of the geometric fluctuations, except in the trivial, uniform case, $`J_A=J_B`$, $`D_A=D_B`$. As in the case of the spin-$`1/2`$ critical fixed point ( $`D_A=D_B=\mathrm{}`$), there also appears a two-cycle of the recursion relations. This two-cycle is located at $`(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})_1=(0.3759\mathrm{},4.5304\mathrm{},0.1472\mathrm{},8.0027\mathrm{}),`$ and $`(K_A^{},K_B^{},\mathrm{\Delta }_A^{},\mathrm{\Delta }_B^{})_2=(2.0653\mathrm{},0.2772\mathrm{},3.4221\mathrm{},0.0905\mathrm{}).`$ As discussed in a previous publication , we should now study the behaviour of the second iterate of the recursion relations around any one of the points belonging to the two-cycle. From the linearization of the second iterates of the recursion relations about these points, we have the eigenvalues $`\mathrm{\Lambda }_1=114.3038\mathrm{},\mathrm{\Lambda }_2=82.0353\mathrm{},\mathrm{\Lambda }_3=5.1867\mathrm{},\mathrm{\Lambda }_4=0.0014\mathrm{}.`$ Now there is at least one eigenvalue with modulus less than $`1`$, which guarantees that the two-cycle is physically accessible. From a numerical analysis of the connectivity of the flow lines in parameter space, we can check that this two-cycle is indeed associated with the tricritical behaviour (in analogy with the second-order transitions associated with the two-cycle in the spin-$`1/2`$ subspace). For $`m=27`$, corresponding to $`d=4`$, we have the same general features. The uniform tricritical fixed point is fully unstable, and there appears a two-cycle, which is presumably associated with a novel tricritical behaviour. Numerical calculations point out that the two-cycle appears as a kind of bifurcation of the uniform tricritical fixed point, at $`m3.45`$ (corresponding to $`d2.13`$). Further numerical work shows that, for a given ratio $`J_A/J_B`$, the temperature that locates the system inside the basin of attraction of the two-cycle (that is, the tricritical temperature) is systematically lower as compared with the tricritical temperature of the uniform model (and thus leads to a smaller first-order region in the phase diagram). These features are also generally present in disordered tricritical systems in three dimensions . For the aperiodic rule $`AAAB`$, $`BAAA`$, with a smaller wandering exponent ($`\omega =0`$ in comparison with $`\omega =\mathrm{ln}2/\mathrm{ln}3`$ for the previous sequence), we have a different set of recursion relations, but the results are similar. In this case, the calculations show that the uniform tricritical fixed point remains accessible (there is one eigenvalue with modulus less than $`1`$) whatever the value of $`m`$. The smaller eigenvalue around the tricritical fixed point approaches unity as $`m`$ goes to infinity. Weaker geometric fluctuations are therefore unable to relevantly perturb the system, whose multicritical behaviour remains unchanged with respect to the uniform model. In conclusion, we have studied the effects of deterministic aperiodicity on a simple system that displays first and second-order transition lines and a tricritical point. Using a simple Migdal-Kadanoff approximation, we show that a certain class of non-random geometric fluctuations may change the tricritical behaviour (by turning into a fully unstable node the fixed point associated with the uniform tricritical behaviour). There is then a two-cycle of the recursion relations that is shown to control the tricritical behaviour. Numerical calculations indicate a depression of the tricritical temperatures (as it used to happen in the presence of quenched disorder). In spite of the limitations of the MK approximation (or, alternatively, the artificiality of the hierarchical lattices in which it turns out to be exact), the results of this investigation provide suggestions for the analysis of multicritical behaviour in similar systems on realistic Bravais lattices. The authors wish to thank A. P. Vieira for useful discussions. This work has been supported by the Brazilian agency FAPESP.
warning/0004/cond-mat0004045.html
ar5iv
text
# Valence Bond Ground States in Quantum Antiferromagnets and Quadratic Algebras ## I Introduction Quadratic algebras and their representations have being extensively used recently in order to study the probability distributions of steady states of one-dimensional stochastic processes with open boundaries or on a ring . The basic idea is that if the Hamiltonian of a quantum chain which gives the time evolution of the system, has eigenvalue zero, the ket wave functions which are related to the steady states probability distributions have a simple expression in terms of a certain quadratic algebra determined by the bulk rates. This algebra has representations fixed by the boundary conditions, the corresponding matrices act in an auxiliary vector space. All correlation functions can be computed from these ket wave functions. The aim of this paper is to ”import” these techniques to equilibrium statistical physics and stress the limitations and differences. For stochastic processes the lowest eigenvalue of the Hamiltonian which gives the time evolution of the system is zero. This is not the case for most of the Hamiltonians which are interesting in equilibrium problems. Therefore the possible applications of the algebraic approach to ground-states is bound to be more limited. Another difference is that in equilibrium and periodic boundary conditions, the ground-state can have momentum non-zero (is not translational invariant). This can’t be the case for stochastic processes since the components of the ground-state ket vector have to be positive numbers (they are probabilities). Another difference appears when we want to calculate correlation functions which are expressed in terms of vacuum expectation values (implying the bra AND ket vacua). As we are going to see the expressions of the correlation functions are very similar in both cases. Actually the quadratic algebras approach was implicit already used in equilibrium problems where it is known as the matrix product approach . The matrices used are in fact representations of certain algebras. We hope to convince the reader that the algebraic approach is not only more aesthetic but more powerful since it makes contact with known results obtained in mathematics. Finally, we would like to mention that matrix product approach has been used as an alternative to the density matrix renormalization group method . How the methods presented in this paper can be applied to this problem is an open question. The application of quadratic algebras to zero-energy states is presented in Sec.2. Much of the contend of this Section is already known. What is new is how to handle zone boundary states which have momentum $`\pi `$. In Sec.3 we give an application. The idea is simple: in the study of quantum groups , in order to find the non-commutative manifold in which they act, Reshetikhin et al have introduced projector operators out of which one can build quantum chains having the quantum algebra as symmetry. Moreover, one gets for free a quadratic algebra (the manifold of the quantum group) which can be used to write the zero energy eigenfunctions of quantum chains build using the projector operators. These chains are not exactly integrable in the Yang-Baxter sense. We have considered, as an example, the $`O(N)`$ case for which we get an $`N`$-state Hamiltonian. The quadratic algebra turns out to be the $`q`$-deformed Clifford algebra. In the special case $`N=3`$ and $`q=1`$ one recovers the model with valence bond ground state (VBS) of Affleck et al . (The q-deformed case can be found in Refs. ). The $`N=4`$ case is discussed in the Appendix A, it is a special case of the extended Hubbard model . The Hamiltonians we consider can be mapped into quantum spin ladder models and find applications in this context. We are going to show that for periodic boundary conditions and an even number of sites, we find a unique momentum zero ground-state. For $`N`$ even, we also find one zone boundary state. For free boundary conditions, we find $`2^{N1}`$ ground-states. This degeneracy can be lifted adding boundary fields. In Sec. 4 we show how to choose the boundary conditions in order to get an unique vacuum. The boundary terms break the symmetry of the quantum chain. The calculation of all the correlation lengths (for any $`N`$) is presented in Sec.5. It is shown that for large $`N`$ the correlation lengths diverge. In Appendix B we show how to compute the correlation function for some parity violating operators, appearing in the case where $`N`$ is even. This problem is interesting in the case of periodic boundary conditions when the ground-state is twice degenerate even for a finite number of sites. The conclusions can be found in Sec.6. ## II Zero energy states and quadratic algebras. The application of quadratic algebras to the zero energy ket wave functions for diffusion-reaction processes is well known , in this section we will do a trivial extension to equilibrium processes and show how to compute correlation functions. We consider a most general one-dimensional quantum chain with $`N`$ states, $`L`$ sites and nearest-neighbour two body interactions. The Hamiltonian is: $$H=\underset{k=1}{\overset{L1}{}}H_k++.$$ (1) The bulk terms ($`k=1,\mathrm{},L1`$) and the left and right boundary terms are: $$H_k=\underset{\alpha ,\beta ,\gamma ,\delta =1}{\overset{N}{}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }E_k^{\gamma \alpha }E_{k+1}^{\delta \beta }$$ (2) $$=\underset{\alpha ,\beta =1}{\overset{N}{}}L_\beta ^\alpha E_1^{\beta \alpha },=\underset{\alpha ,\beta =1}{\overset{N}{}}R_\beta ^\alpha E_L^{\beta \alpha }.$$ (3) Here $`E_k^{\alpha \beta }`$ are a basis for $`N\times N`$ matrices on the $`k`$-th site: $$(E^{\alpha \beta })_{\gamma \delta }=\delta _{\alpha \gamma }\delta _{\beta \delta }(\alpha ,\beta ,\gamma ,\delta =1,\mathrm{},N).$$ (4) We will assume that $`H`$ has at least one eigenstate of energy zero $$H|0>=0<0|H=0.$$ (5) Our aim is to describe the bra $`<0|`$ and ket $`|0>`$ states in a simple way. In order to do that, we consider two associative algebras defined by the bulk interaction: $$\underset{\alpha ,\beta =1}{\overset{N}{}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }x_\alpha x_\beta =x_\gamma X_\delta X_\gamma x_\delta $$ (6) $$\underset{\gamma ,\delta =1}{\overset{N}{}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }y_\gamma y_\delta =y_\alpha Y_\beta Y_\alpha y_\beta .$$ (7) If the bulk part of the Hamiltonian is not symmetric, the two algebras are different. Each algebra has $`2N`$ generators $`x_\alpha ,X_\alpha `$ and $`y_\alpha `$, $`Y_\alpha `$,( $`\alpha =1,\mathrm{},N`$), respectively. We define two Fock-like representations of the two algebras: $$<V_K|(X_\alpha \underset{\beta =1}{\overset{N}{}}L_\alpha ^\beta x_\beta )=0(X_\alpha +\underset{\beta =1}{\overset{N}{}}R_\alpha ^\beta x_\beta )|W_K>=0$$ (8) $$<V_B|(Y_\beta \underset{\alpha =1}{\overset{N}{}}L_\alpha ^\beta y_\alpha )=0(Y_\beta +\underset{\alpha =1}{\overset{N}{}}R_\alpha ^\beta y_\alpha )|W_B>=0.$$ (9) Here $`<V_K|,|W_K>,<V_B|`$ and $`|W_B>`$ are the bra and ket reference states defined by the equations (8) and (9) in AUXILIARY spaces. We make now the connexion between the two algebras and the zero energy eigenstates of the Hamiltonian. The basis in the ket vector space in which the Hamiltonian acts is: $$u_{\alpha _1}u_{\alpha _2}\mathrm{}u_{\alpha _L}(\alpha _k=1,2,\mathrm{},N)$$ (10) the $`N`$-dimensional vector $`u_{\alpha _k}`$ is in the $`k`$-th site and has the component $`\alpha _k`$ equal to one and the others zero: $$(u_{\alpha _k})_\beta =\delta _{\alpha _k,\beta }(\beta =1,2,\mathrm{},N).$$ (11) We denote the basis in the bra vector space in which the Hamiltonian acts by $$u_{\alpha _1}^Tu_{\alpha _2}^T\mathrm{}u_{\alpha _L}^T.$$ (12) The scalar product is obviously $$<u_{\alpha _k}^Tu_{\beta _k}>=\delta _{\alpha _k,\beta _k}.$$ (13) One can prove that the unnormalized bra and ket vacua can be written using the two quadratic algebras: $$|0>=\underset{\alpha _1,\mathrm{},\alpha _L=1}{\overset{N}{}}<V_K|x_{\alpha _1}\mathrm{}x_{\alpha _L}|W_K>u_{\alpha _1}\mathrm{}u_{\alpha _L}$$ (14) $$<0|=\underset{\alpha _1,\mathrm{},\alpha _L=1}{\overset{N}{}}<V_B|y_{\alpha _1}\mathrm{}y_{\alpha _L}|W_B>u_{\alpha _1}^T\mathrm{}u_{\alpha _L}^T.$$ (15) Notice that the generators $`X_\alpha `$ and $`Y_\alpha `$ don’t appear in the expressions of the wave functions. One can also show that the quadratic algebras exist, and that one can find representations satisfying the conditions (8) and (9). Moreover, one can show that all the zero energy wave functions can be obtained in this way . In the case of periodic boundary conditions, and translationally invariant zero energy eigenfunctions, one can use the expressions (14) and (15) making the substitution: $$<V_K|\mathrm{}|W_K>\text{Tr}(\mathrm{})<V_B|\mathrm{}|W_B>\text{Tr}(\mathrm{})$$ (16) provided that the algebra has a trace operation. As opposed to the case of the Hamiltonian with open boundaries, for periodic boundary conditions, it is not clear in which cases one obtains in this way all the zero energy eigenfunctions. A simple counter-example was given in Ref. in which it is shown that there are zero energy eigenfunctions which can’t be obtained using the algebraic method given by equation (16). On the other hand, examples are known where indeed all the eigenfunctions are obtained. Ground-state wave functions can correspond to zone boundary states (momentum $`\pi `$). One can show that if the algebra (6) has the Str operation with the properties: $`\text{Str}(x_{\alpha _1}x_{\alpha _2}\mathrm{}x_{\alpha _L})`$ $`=`$ $`\text{Str}(x_{\alpha _L}x_{\alpha _1}x_{\alpha _2}\mathrm{}x_{\alpha _{L1}})`$ (17) $`\text{Str}(X_{\alpha _1}x_{\alpha _2}\mathrm{}x_{\alpha _L})`$ $`=`$ $`\text{Str}(x_{\alpha _L}X_{\alpha _1}x_{\alpha _2}\mathrm{}x_{\alpha _{L1}})`$ (18) than the ket vector $$|0>=\underset{\alpha _1,\mathrm{},\alpha _L=1}{\overset{L}{}}\text{Str}(x_{\alpha _1}\mathrm{}x_{\alpha _L})u_{\alpha _1}\mathrm{}u_{\alpha _L}$$ (19) satisfies equation (5) and it is obviously a zone boundary state. Similar expressions can be used for the algebra (7) and the bra eigenvector. The Str (called supertrace) operation is taken from the theory of superalgebras and it implies that the $`x_\alpha `$ and $`X_\alpha `$ are odd generators in this algebra. In particular if in the algebra (6) one takes $`X_\alpha `$ c-numbers (this is often done for diffusion processes ), the algebra can’t have the Str operation. In Sec.3 we will show in examples how the Str operation works. As for translationally invariant ground-states it is not known if all of the zone boundary states can be obtained using equation (19). Before showing how to compute correlation functions, let us see what are the consequences for the quadratic algebras of the existence of a symmetry of the Hamiltonian. Let us assume that the operator: $$A=\underset{k=1}{\overset{L}{}}\underset{\mu ,\nu =1}{\overset{N}{}}A_{\mu \nu }E_k^{\mu \nu }$$ (20) commutes with the bulk part of the Hamiltonian, i. e., $$[A,\underset{k=1}{\overset{L1}{}}H_k]=0.$$ (21) Simple arithmetics gives the relations: $`{\displaystyle \underset{\alpha ,\beta =1}{\overset{N}{}}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }[(A_{\alpha \mu }x_\mu )x_\beta +x_\alpha (A_{\beta \mu }x_\mu )]`$ $`=`$ (22) $`(A_{\gamma \mu }x_\mu )X_\delta +x_\gamma (A_{\delta \mu }X_\mu )`$ $``$ $`(A_{\gamma \mu }X_\mu )x_\delta X_\gamma (A_{\delta \mu }x_\mu ).`$ (23) This relation gives a set of simplified algebraic relations among the generators of the algebra and at the same time, shows that the generators are tensor operators. (A relation similar to (22) can be obtained for the generators $`y_\alpha `$ and $`Y_\alpha `$). As an example, let us choose $`A_{11}=1`$ and all the other matrix elements zero in (20). Using (22) one obtains: $$\underset{\alpha =1}{\overset{N}{}}(\mathrm{\Gamma }_{\gamma \delta }^{1\alpha }x_1x_\alpha +\mathrm{\Gamma }_{\gamma \delta }^{\alpha 1}x_\alpha x_1)=\delta _{\gamma ,1}(x_1X_\delta X_1x_\delta )+\delta _{\delta 1}(x_\gamma X_1X_\gamma x_1).$$ (24) Similar relations can be obtained in the case of quantum algebra symmetries when the operator A has not the simple expression (20). We now show how to compute a two-point function. This calculation is interesting when the ground-state energy is zero. Consider two local operators $`P_r`$ and $`Q_s`$ on the $`r`$ and $`s`$ sites. They act on the basis (10) as follows: $$P_ru_{\alpha _r}=\underset{\beta _r=1}{\overset{N}{}}P_{\beta _r,\alpha _r}u_{\beta _r};Q_su_{\alpha _s}=\underset{\beta _s=1}{\overset{N}{}}Q_{\beta _s,\alpha _s}u_{\beta _s}.$$ (25) We want to compute the expression: $$G_{r,s}=\frac{<0|P_rQ_s|0>}{Z}$$ (26) where $`<0|`$ and $`|0>`$ are given by equations (14) and (15) and $`Z`$ is a normalization factor coming from the fact that (14) and (15) give unnormalized wave functions. It is useful to define the following quantities (all related to the auxiliary space) $$C=\underset{\alpha =1}{\overset{N}{}}x_\alpha y_\alpha $$ (27) $$P=\underset{\alpha ,\beta =1}{\overset{N}{}}P_{\alpha \beta }x_\beta y_\alpha ,Q=\underset{\alpha ,\beta =1}{\overset{N}{}}Q_{\alpha ,\beta }x_\beta y_\alpha $$ (28) and $$<V_B|<V_K|=<V|;|W>=|W_K>|W_B>.$$ (29) Using equations (27)-(29), the two-point function (24) has the following simple expression: $$G_{r,s}=\frac{1}{Z}<V|C^{r1}PC^{sr1}QC^{Ls}|W>$$ (30) where $$Z=<V|C^L|W>.$$ (31) Notice that $`C`$ plays the role of a space evolution operator in the auxiliary space but the analogy with a quantum mechanical problem can’t be pushed further since $`<V|`$ and $`|W>`$ are not eigenfunctions of C. Nevertheless one can see that the spectrum of C gives all the correlation lengths. For periodic boundary conditions, we have to make the following substitution: $`<V|\mathrm{}|W>\text{Tr}(\mathrm{})`$ (32) $`<V|\mathrm{}|W>\text{Str}(\mathrm{})`$ (33) for translationally invariant states, or for zone boundary states, in equations (30) and (31). Let us observe that the expressions (27)-(31) are similar to the ones one obtain for stochastic processes . The difference is that instead of dealing with only one algebra (given by equation (6)), one has the tensor product of two algebras. If the algebra (7) has a one-dimensional representation, (this is always the case for diffusion processes with exclusion for example ), the correlation functions computed using the ket vector only or the bra and ket vector (vacuum expectation values), coincide. Expressions like (27)-(31) have been used in a different context in the matrix product approach to the density matrix renormalization group method . In this case the $`x_\alpha `$ are matrices obtained using the variational method and not using quadratic algebras defined by the Hamiltonian using equation (6). Besides they have to satisfy the condition: $$\underset{\alpha =1}{\overset{N}{}}x_\alpha x_\alpha ^+=1.$$ (34) As we are going to see in the next section, this condition is not necessarily fulfilled in our applications. ## III $`q`$-deformed $`O(N)`$ symmetric, $`N`$-state quantum chains. The quantum chains describing stochastic processes are given by non-hermitian Hamiltonians which always have zero as lowest eigenvalue. The quadratic algebra always exists and the problem is to find representations of the algebra. In equilibrium problems one is interested in hermitian Hamiltonians which in general don’t have zero as the lowest eigenvalue and therefore one has to find Hamiltonians which have this property. In order to illustrate the method, in this paper we have chosen an easy way: using known results in the theory of quantum groups. In this way we get not only hermitian quantum chains which have zero for the ground state energy but also quadratic algebras with known representations. ### A The bulk Hamiltonian. Reading the paper of Reshetikhin et al one can notice that there are several expressions of the form (6) with the $`X_{\alpha _s}`$ equal to zero. We will choose the one where $`\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }`$ are projector operators of rank $`N(N+1)/21`$ for the $`q`$-deformed $`B(n)`$ series ($`N=2n+1`$) and $`D(n)`$ series ($`N=2n`$). The $`x_\alpha `$ are the generators of the non-commutative algebra of the manifold where the quantum groups act. Similar expressions for the $`Sp(n)`$ and $`Osp(m/n)`$ algebras and superalgebras can also be obtained . In the present paper, we confine ourselves to the q-deformed $`O(N)`$ case. As we will show we will use these projectors in order to write Hamiltonians for quantum chains. The projector operators have the following expressions: $`P_k^{(+)}={\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta =1}{\overset{N}{}}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }E_k^{\gamma \alpha }E_{k+1}^{\delta \beta }={\displaystyle \frac{1}{q+q^1}}[q{\displaystyle \underset{\alpha \alpha ^{}}{}}E_k^{\alpha \alpha }E_{k+1}^{\alpha \alpha }+(qq^1){\displaystyle \underset{\alpha >\beta }{}}E_k^{\beta \beta }E_{k+1}^{\alpha \alpha }`$ (35) $`+\delta _{N,2n+1}E_k^{\frac{N+1}{2}\frac{N+1}{2}}E_{k+1}^{\frac{N+1}{2}\frac{N+1}{2}}+q^1{\displaystyle \underset{\alpha ,\beta =1}{\overset{N}{}}}E_k^{\alpha \alpha }E_{k+1}^{\beta \beta }+{\displaystyle \underset{\alpha \beta ,\beta ^{}}{}}E_k^{\beta \alpha }E_{k+1}^{\alpha \beta }`$ (36) $`+q^1{\displaystyle \underset{\alpha \alpha ^{}}{}}E_k^{\alpha \alpha ^{}}E_{k+1}^{\alpha ^{}\alpha }{\displaystyle \frac{q^{\frac{N}{2}}}{[\frac{N}{2}]_q}}{\displaystyle \underset{\alpha ,\beta =1}{\overset{N}{}}}E_k^{\alpha ^{}\beta }E_{k+1}^{\alpha \beta ^{}}q^{\rho _\alpha \rho _\beta }(qq^1){\displaystyle \underset{\alpha >\beta }{}}E_k^{\alpha ^{}\beta }E_{k+1}^{\alpha \beta ^{}}q^{\rho _\alpha \rho _\beta }]`$ (37) where $`q`$ is a deformation parameter (taken real in this paper) and we use the notation $$[n]_q=\frac{q^nq^n}{qq^1}\text{and}\alpha ^{}=N+1\alpha (\alpha =1,\mathrm{},N).$$ (38) In equation (35) we also denote $$(\rho _1,\mathrm{},\rho _N)=(n\frac{1}{2},n\frac{3}{2},\mathrm{},\frac{1}{2},0,\frac{1}{2},\mathrm{},n+\frac{1}{2}),$$ (39) for $`N=2n+1`$, and $$(\rho _1,\mathrm{},\rho _N)=(n1,n2,\mathrm{},1,0,0,1,\mathrm{},n+1)$$ (40) for N=2n. By definition we have $$(P_k^{(+)})^2=P_k^{(+)}.$$ (41) Since the matrix $`\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }`$ in (35) is symmetric, i. e., $`\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }=\mathrm{\Gamma }_{\alpha \beta }^{\gamma \delta }`$, the two associated algebras to the projector (35): $$\underset{\alpha ,\beta =1}{\overset{N}{}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }x_\alpha x_\beta =0\text{and}\underset{\alpha ,\beta =1}{\overset{N}{}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }y_\alpha y_\beta =0$$ (42) are identical and therefore we give only one of them. It is convenient to denote (for obvious reasons); for N=2n: $$x_1=a_n,x_2=a_{n1},\mathrm{},x_n=a_1,x_{n+1}=a_1^+,x_{n+2}=a_2^+,\mathrm{},x_{2n}=a_n^+$$ (43) and for, N=2n+1 $$x_1=a_n,x_2=a_{n1},\mathrm{},x_n=a_1,x_{n+1}=\frac{1}{\sqrt{s+s^1}}\mathrm{\Sigma },x_{n+2}=a_1^+,\mathrm{},x_{2n+1}=a_n^+$$ (44) where $`s=\sqrt{q}`$. The $`2n`$ $`q`$-deformed fermionic creation and annihilation operators $`a_\alpha ,a_\alpha ^+`$ and the ”$`\gamma ^5`$”-type generator $`\mathrm{\Sigma }`$ satisfy the following relations: $`qa_\beta a_\alpha +a_\alpha a_\beta `$ $`=`$ $`0(\beta >\alpha )`$ (45) $`qa_\beta a_\alpha ^++a_\alpha ^+a_\beta `$ $`=`$ $`0(\beta >\alpha )`$ (46) $`\mathrm{\Sigma }a_\alpha +qa_\alpha \mathrm{\Sigma }`$ $`=`$ $`0,\mathrm{\Sigma }^+=\mathrm{\Sigma }`$ (47) $`a_\alpha a_\alpha ^++a_\alpha ^+a_\alpha `$ $`=`$ $`qa_{\alpha +1}a_{\alpha +1}^++q^1a_{\alpha +1}^+a_{\alpha +1}(1\alpha n1)`$ (48) $`qa_1a_1^++q^1a_1^+a_1`$ $`=`$ $`\mathrm{\Sigma }^2.`$ (49) The above algebra has a central element : $$\zeta =a_na_n^++a_n^+a_n$$ (50) and an obvious representation is: $`a_k`$ $`=`$ $`11\mathrm{}as^{\sigma ^z}\sigma ^zs^{\sigma ^z}\sigma ^z\mathrm{}s^{\sigma ^z}\sigma ^z,(k=1,\mathrm{},n)`$ (51) $`\mathrm{\Sigma }`$ $`=`$ $`s^{\sigma ^z}\sigma ^zs^{\sigma ^s}\sigma ^z\mathrm{}s^{\sigma ^z}\sigma ^z,`$ (52) with $`a=\left(\begin{array}{cc}\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}& \\ \mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \end{array}\right),a^+=\left(\begin{array}{cc}\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ \mathrm{1\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \end{array}\right),\sigma ^z=\left(\begin{array}{cc}\mathrm{1\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ 01& \end{array}\right).`$ (59) In the first line of (51) the operator $`a`$ is in the $`k`$-th position, and the operator $`s^{\sigma ^z}\sigma ^z`$ appears in the positions $`k+1,\mathrm{},n`$. The fact that the algebra has finite-dimensional representations, makes all calculations much simpler (see Secs. 5 and 6) as compared with the cases when the algebra has infinite dimensional representations. Notice also that for $`q1`$, the generators $`x_\alpha `$ do not satisfy the relation (34). We make now the connexion between the projectors (35) and the quantum chain (1). Since the lowest eigenvalue $`E`$ of a projector operator is zero, we can choose in (1): $$H_k=P_k^{(+)}$$ (60) Notice that $`H_k`$ is hermitian and therefore since as mentioned its lowest eigenvalue is zero, $`H`$ for periodic or free boundary conditions has also zero as its lowest eigenvalue. The problem of other boundary conditions is going to be discussed in Sec.4. ### B Ground-states for periodic and free boundary conditions. We start with periodic boundary conditions. We first consider zero momentum states. Using equation (14) (together with the substitution given by (16) as well as the representation (51) we get for all $`N`$, one single ket vector of energy zero for $`L`$ even and none for $`L`$ odd. This result is confirmed by the spectra obtained from the numerical diagonalization of several Hamiltonians (various $`L`$ and $`N`$). This check was necessary since as mentioned in Sec.2 there is no theorem which assures us that there are no zero energy eigenfunctions which are not obtained using the algebraic procedure. We now look for zone boundary states and therefore look for a definition of the Str operation such that the relations (17) and (19) are satisfied. We consider the matrix $`J`$ defined by $$J=\sigma ^z\sigma ^z\mathrm{}\sigma ^z.$$ (61) A vector $$|v>=|v_1>|v_2>\mathrm{}|v_n>$$ (62) is called even (odd) if it is an eigenvector of $`J`$ corresponding the the eigenvalue +1 respectively -1. A matrix is called even if it takes an even (odd) vector into an even (odd) vector. A matrix is called odd if it takes an even (odd) vector into an odd (even) vector. For example, the matrices $`a_k`$ in equation (51) are odd but the matrix $`\mathrm{\Sigma }`$ is even. Consider now the matrix $$A=A_1A_2\mathrm{}A_n.$$ (63) We define $$\text{Str}(M)=\text{Tr}(JM).$$ (64) It is easy to check that if $`A`$ and $`B`$ are odd matrices, than $$\text{Str}(AB)=\text{Str}(BA).$$ (65) If one of the two matrices is even and the other one is odd $$\text{Str}(AB)=0.$$ (66) If the two matrices $`A`$ and $`B`$ are even $$\text{Str}(AB)=\text{Str}(BA).$$ (67) From this properties we learn that in order to satisfy the relations (17) (the relations in the second line of (19) are automatically satisfied since $`X_\alpha =0`$), the $`x_\alpha `$’s have to be all odd generators. This excludes the case of $`N=2n+1`$ because of the appearance of the sigma generator which is even. For $`L`$ odd and $`N=2n`$ all the supertraces are zero and again we can’t obtain a boundary state which is physically correct. For $`N`$ and $`L`$ even we expect therefore a unique zone boundary state. This is what is also seen in numerical diagonalizations for all $`L`$ except for $`q=1`$ and small values of $`L`$ where something subtle happens. We illustrate the phenomenon taking $`N=4`$. Using equations (51),(61) and (64) we obtain: $$\text{Str}(a_2^+a_2)=0,\text{Str}(a_1^+a_1)=qq^1,$$ (68) which would imply that for $`q=1`$ and $`L=2`$ there are no zone boundary states. Actually there are two of them which can be obtained taking instead of $`J`$ given by equation (61), two alternative expressions: $$1\sigma ^z\text{or}\sigma ^z1.$$ (69) These expressions can’t be used however for monomials with more than two generators (the property (17) is not valid anymore). In the spectra for periodic boundary conditions as seen in numerical diagonalizations, there are no zero energy states besides those mentioned above. The existence for $`N`$ even of a degenerate ground state, one of positive parity (momentum zero, obtained with the help of the Tr operation) that we denote by $`|0,+>`$ and one of negative parity (momentum $`\pi `$, obtained with the help of the Str operation that we denote by $`|0,>`$), allows for the existence of correlation functions of operators which break parity. For example one of the operators $`P`$ or $`Q`$ in equation (27) can break parity. In the Appendix B we show how to compute the correlation functions for this case (one considers matrix elements $`<0,|\mathrm{}|0,+>`$ for example). A somehow similar problem occurs in spontaneously dimerized spin ladders . We would like to stress that in our case the degeneracy of the vacuum takes place even for finite number of sites. We now consider free boundary conditions. An inspection of equation (8) shows that it brings no constrains therefore instead of equation (14) we have: $$|0>=\underset{\alpha _1,\mathrm{},\alpha _L=1}{\overset{N}{}}x_{\alpha _1}\mathrm{}x_{\alpha _L}u_{\alpha _1}\mathrm{}u_{\alpha _L},$$ (70) where the various independent monomials (words) in the algebra are regarded as a basis in a vector space. Each component of $`|0>`$ in this basis gives a zero energy eigenfunction. Therefore for both $`L`$ even and odd we get $`2^{N1}`$ states. This result was obtained counting the independent words. For small values of $`L`$, the degeneracy can be smaller since higher degree monomials might not yet have appeared. For example for $`N=4`$, and $`L=2`$ the degeneracy is 7 instead of 8 but for $`L=3`$ one obtains already 8. ## IV Boundary conditions compatible with the quadratic algebras The boundary matrices $``$ and $``$ (we will choose them hermitian) have not only to be compatible with the quadratic algebra (see below), but have also to leave the value zero as the lowest eigenvalue. This property is warrantied if the lowest eigenvalues $`E_L`$ and $`E_R`$ are also zero. This follows from the relation: $$E_HE_L+(L1)E+E_R.$$ (71) where $`E_H`$ and $`E`$ are the lowest eigenvalues of $`H`$ and $`H_k`$. Since for the $`q`$-deformed $`O(N)`$ symmetric quantum chains defined in the last section, the algebras (6) and (7) with $`X_\alpha `$ and $`Y_\beta `$ equal to zero are identical, we have to find the matrices $``$ and $``$ as well as the vacua of the auxiliary spaces such that the following relations (obtained from equations (8)-(9)) are satisfied: $$\underset{\beta =1}{\overset{N}{}}R_\alpha ^\beta x_\beta |W_K>=0;\underset{\alpha =1}{\overset{N}{}}R_\alpha ^\beta x_\alpha |W_B>=0,$$ (72) $$\underset{\beta =1}{\overset{N}{}}L_\beta ^\alpha <V_K|x_\beta =0,\underset{\alpha =1}{\overset{N}{}}L_\beta ^\alpha <V_B|x_\alpha =0.$$ (73) We have taken the same representation for the two sets of Clifford generators $`x_\alpha `$ and $`y_\alpha `$. We now show that the solutions of equations (73) can be obtained from those of equations (72). We take the transpose of the two equations (73): $$\underset{\beta =1}{\overset{N}{}}L_\beta ^\alpha x_\beta ^T|V_K^T>=0,\underset{\alpha =1}{\overset{N}{}}L_\beta ^\alpha x_\alpha ^T|V_B^T>=0,$$ (74) and since from equations (43)-(44), we have $`x_\alpha ^T=x_\alpha ^{}=x_{N+1\alpha }`$, we can rewrite the equations (74) as follows: $$\underset{\beta =1}{\overset{N}{}}L_\beta ^{}^\alpha x_\beta |V_K^T>=0,\underset{\alpha =1}{\overset{N}{}}L_\beta ^\alpha ^{}x_\alpha |V_B^T>=0.$$ (75) We can compare now the equations (72) and (75) and deduce that for any solution $`R_\alpha ^\beta `$ of (72) (there are many of them) one gets a solution for $`L_\alpha ^\beta `$ : $$L_\beta ^\alpha =R_\beta ^{}^\alpha ^{}.$$ (76) One can use of course one solution of equations (72) for $`R_\alpha ^\beta `$ and another solution to get $`L_\alpha ^\beta `$ using equation (76). One can easily show that the solutions of (72) have a factorized form: $$R_\beta ^\alpha =re_\alpha f_\beta .$$ (77) It is convenient to choose the following basis in the auxiliary vector-spaces in which $`x_\alpha `$ and $`y_\alpha `$ (replaced formally by $`x_\alpha `$) act (see equations (72) and (51)-(59)): $`|W_K>=({\displaystyle \underset{i=1}{\overset{L}{}}}{\displaystyle \frac{1}{\sqrt{1+\eta _i^2}}})\left(\begin{array}{c}\eta _1\\ 1\end{array}\right)\left(\begin{array}{c}\eta _2\\ 1\end{array}\right)\mathrm{}\left(\begin{array}{c}\eta _n\\ 1\end{array}\right)`$ (84) $`|W_B>=({\displaystyle \underset{i=1}{\overset{L}{}}}{\displaystyle \frac{1}{\sqrt{1+\stackrel{~}{\eta }_i^2}}})\left(\begin{array}{c}\stackrel{~}{\eta }_1\\ 1\end{array}\right)\left(\begin{array}{c}\stackrel{~}{\eta }_2\\ 1\end{array}\right)\mathrm{}\left(\begin{array}{c}\stackrel{~}{\eta }_n\\ 1\end{array}\right).`$ (91) (14)-(15) We are going to consider separately the cases $`N=2,3,\mathrm{},6`$ in order to illustrate the structure of the solutions. As we are going to show for $`N=2`$, the values of $`\eta _1`$ and $`\stackrel{~}{\eta }_1`$ are fixed and besides a common factor, the matrix elements of $``$ contain no parameters. For $`N=3`$ and 4, the parameters $`\eta _1,\eta _2`$ respectively $`\stackrel{~}{\eta }_1,\stackrel{~}{\eta }_2`$ are free and $``$ is given by the parameters of the wave functions (84)-(91) and a common factor. For $`N=5`$ and 6, a new phenomenon appears. The wave function (84) is given by the free parameters $`\eta _1,\eta _2,\eta _3`$, and the wave function (91) is specified by the corresponding parameters $`\stackrel{~}{\eta }_1,\stackrel{~}{\eta }_2,\stackrel{~}{\eta }_3`$. $``$ depends now not only on the parameters of the wave functions but on supplementary free parameters. This implies that different boundary conditions are compatible with the same wave functions (14)-(15). We now consider the boundary conditions for some values of $`N`$. N=2 Since this is a very simple (and trivial) case, we discuss it in detail. From equations (35) and (60) we get: $$H_k=\frac{1}{2}(\sigma _k^z\sigma _{k+1}^z+1).$$ (92) This implies that for free boundary conditions the ground-state is twice degenerate, with antiferromagnetic ordering. This degeneracy is a consequence of the existence of two independent words in the $`O(2)`$ algebra: $`x_1x_2\mathrm{}x_1x_2`$ and $`x_2x_1\mathrm{}x_2x_1`$. Demanding that $``$ is diagonalizable, equations (72) have two solutions: $`=rA;|W_K>=\left(\begin{array}{c}1\\ 0\end{array}\right),|W_B>=\left(\begin{array}{c}1\\ 0\end{array}\right)`$ (97) and $`=rB;|W_K>=\left(\begin{array}{c}0\\ 1\end{array}\right),|W_B>=\left(\begin{array}{c}0\\ 1\end{array}\right)`$ (102) where $`A=\left(\begin{array}{cc}\mathrm{1\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ \mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \end{array}\right),B=\left(\begin{array}{cc}\mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}& \\ \mathrm{0\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}& \end{array}\right)`$ (107) and $`r>0`$. We consider separately the two cases: a) $`=rA`$, $`=lA`$ with $`r,l>0`$. For lattice size $`L`$ even, we obtain no zero energy eigenstate. The matrix elements (14)-(15) vanish. For $`L`$ odd, one obtains an unique zero energy ground-state. b) $`=rA,=lB`$ For $`L`$ even, one obtains an unique zero energy ground-state and none for $`L`$ odd. N=3 The solutions are : $`f_1=s\eta _1,f_2=\sqrt{s+s^1},f_3=(s\eta _1)^1`$ (108) $`e_1=s\stackrel{~}{\eta }_1,e_2=\sqrt{s+s^1},e_3=(s\stackrel{~}{\eta }_1)^1`$ (109) and the eigenvalues of $``$ are: zero two times and $$r\left[1+\frac{1}{s+s^1}(s^2\eta _1\stackrel{~}{\eta }_1+(s^2\eta _1\stackrel{~}{\eta }_1)^1)\right]$$ (110) this relation imposes $`r>0`$. Since we want $``$ symmetric, we take (see equation (77)) $`\eta _1=\stackrel{~}{\eta }_1`$ and therefore: $$e_\alpha =f_\alpha (\alpha =1,2,\mathrm{},N).$$ (111) N=4 One gets: $$f_1=s\eta _2,f_2=\eta _1,f_3=\eta _1^1,f_4=(s\eta _2)^1.$$ (112) The matrix $``$ has three eigenvalues zero and one equal to: $$r[(s\eta _2)^2+\eta _1^2+\eta _1^2+(s\eta _2)^2].$$ (113) Notice that for $`N=3`$ and 4, the parameters of the vacua and $`r`$ determine the $``$ matrix. This is abound to change for larger values of $`N`$. N=5 The solution is $`f_1=\eta _2s\left(1+a{\displaystyle \frac{ss^1}{s+s^1}}\right),f_2=\eta _1(1a),f_3={\displaystyle \frac{2a}{\sqrt{s+s^1}}}`$ (114) $`f_4=\eta _1^1(a+1),f_5=\eta _2^1s^1\left(1+a{\displaystyle \frac{ss^1}{s+s^1}}\right),`$ (115) where $`a`$ ia an additional free parameter. $``$ has now four eigenvalues zero and one equal to $`r_{i=1}^5f_i^2`$. N=6 One gets: $`f_1=\eta _3s\left(a+{\displaystyle \frac{ss^1}{2}}\right),f_2=\eta _2\left(a{\displaystyle \frac{s+s^1}{2}}\right),f_3=\eta _1`$ (116) $`f_4=\eta _1^1,f_5=\eta _2^1\left(a+{\displaystyle \frac{s+s^1}{2}}\right),f_6=s^1\eta _3^1\left(a+{\displaystyle \frac{ss^1}{2}}\right)`$ (117) with $`a`$ arbitrary. $``$ has now five eigenvalues zero and one equal to $`r_{i=1}^6f_i^2`$. Notice that for $`N=5`$ and 6 the $`f`$’s depend not only on $`\eta _1,\eta _2`$ and $`\eta _3`$ but also on the supplementary parameter $`a`$. This implies that the same wave function can be used for different boundary matrices. We also notice that taking $`r`$ positive makes sure that the lowest eigenvalue stays zero. One can obtain $``$ matrices with only non-vanishing element on the diagonal (like in equation (107)) taking one of the $`\eta _i`$ equal to zero or infinity. This remains valid for any $`N`$. For larger values of $`N`$ the number of free parameters increases and it is certainly not our purpose to give here the general solution. We would like to stress that for $`N>2`$, the boundary conditions can break all the symmetries of the Hamiltonian. ## V Diagonalization of the $`C`$ matrix and Calculation of the correlation lengths of the $`q`$-deformed $`O(N)`$ quantum chain. It is necessary to have a new look at the expression (30) of the two-point correlation function. In the last section we have shown how to get the bra and ket vacua ($`<V|`$ and $`|W>`$, respectively) in the auxiliary spaces. In order to proceed further, one has to find the similarity transformation $`S`$ which diagonalizes the matrix $`C`$: $$C=SC_DS^1.$$ (118) The matrix $`C`$ is given by equation (27). In this equation the $`x_\alpha `$ and $`y_\alpha `$ are the generators of the two identical algebras (see (43)-(45)) having the representation (51)-(59). We will consider separately the cases where $`N`$ is even or odd. a) N=2n It is convenient to write $`C`$ as a four-state Hamiltonian with $`n`$ sites in the auxiliary space: $$C^{(N)}=E_1F_2F_3\mathrm{}F_n+E_2F_3F_4\mathrm{}F_n+\mathrm{}+E_n$$ (119) where the matrices $`E_i`$ and $`F_i`$ act on the $`i`$-th site and have the expression: $`\begin{array}{cccc}(1,1)& (1,2)& (2,1)& (2,2)\end{array}`$ (121) $`E=aa=`$ $`\left(\begin{array}{cccc}\mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}1}\\ \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}\\ \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}\\ \mathrm{\hspace{0.33em}\hspace{0.33em}1}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}\hspace{0.33em}0}\end{array}\right)`$ (126) $`\begin{array}{cccc}(1,1)& (1,2)& (2,1)& (2,2)\end{array}`$ (129) $`F=s^{\sigma ^z}\sigma ^zs^{\sigma ^z}\sigma ^z=`$ $`\left(\begin{array}{cccc}q& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& 0\\ \mathrm{\hspace{0.33em}\hspace{0.33em}0}& 1& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& 0\\ \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& 1& 0\\ \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& \mathrm{\hspace{0.33em}\hspace{0.33em}0}& q^1\end{array}\right)`$ (134) where $`a=\left(\begin{array}{cc}0& 1\\ 0& 0\end{array}\right);\sigma ^z=\left(\begin{array}{cc}1& 0\\ 0& 1\end{array}\right).`$ (140) The basis vectors in the tensor products (121)-(129) correspond to the two-dimensional representations used in (51)-(59). In this basis the vacuum $`|W>`$ in the auxiliary space has the expression: $$|W>=V^{(1)}V^{(2)}\mathrm{}V^{(n)}$$ (141) where $$V=\frac{1}{\sqrt{(1+\eta ^2)(1+\stackrel{~}{\eta }^2)}}\left(\begin{array}{c}\eta \\ 1\end{array}\right)\left(\begin{array}{c}\stackrel{~}{\eta }\\ 1\end{array}\right).$$ (142) Notice the recurrence relation: $$C^{(N+2)}=C^{(N)}F_{n+1}+E_{n+1}$$ (143) that we are going to use later on. If $`q=1`$, the diagonalization of $`C`$ is trivial since $`E`$ and $`F`$ commute. Using the similarity transformation $`U={\displaystyle \frac{\sqrt{2}}{2}}\left(\begin{array}{cccc}1& \mathrm{\hspace{0.33em}0}& \mathrm{\hspace{0.33em}0}& 1\\ 0& \sqrt{2}& \mathrm{\hspace{0.33em}0}& \mathrm{\hspace{0.33em}0}\\ 0& \mathrm{\hspace{0.33em}0}& \sqrt{2}& \mathrm{\hspace{0.33em}0}\\ 1& \mathrm{\hspace{0.33em}0}& \mathrm{\hspace{0.33em}0}& \mathrm{\hspace{0.33em}1}\end{array}\right)`$ (148) we have $$E_D=UEU^1=\frac{1}{2}(\sigma ^z1+1\sigma ^z);F_D=U^1FU=\sigma ^z\sigma ^z.$$ (149) It is convenient to write $`C_D`$ as a one-dimensional two-state spin chain with $`2n`$ sites: $`C_D^{(N)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[(\sigma _1^z+\sigma _2^z)(\sigma _3^z\sigma _4^z)\mathrm{}(\sigma _{2n1}^z\sigma _{2n}^z)+`$ (151) $`(\sigma _3^z+\sigma _4^z)(\sigma _5^z\sigma _6^z)\mathrm{}(\sigma _{2n1}^z\sigma _{2n}^z)+\mathrm{}+(\sigma _{2n1}^z+\sigma _{2n}^z)].`$ In order to simplify the expression (151), it is useful to look at $`C`$ as a function defined on the abelian group $`Z_2Z_2\mathrm{}Z_2=(Z_2)^{2n}`$. In order to do so, we write: $$\sigma _k^z=(1)^{ϵ_k}(ϵ_k=0,1).$$ (152) Using this notation, instead of equation (151) we obtain: $`C_D^{(N)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}[(1)^{ϵ_1}+(1)^{ϵ_2})((1)^{ϵ_3}(1)^{ϵ_4})\mathrm{}((1)^{ϵ_{2n1}}(1)^{ϵ_{2n}})+`$ (155) $`((1)^{ϵ_3}+(1)^{ϵ_4})((1)^{ϵ_5}(1)^{ϵ_6})\mathrm{}((1)^{ϵ_{2n1}}(1)^{ϵ_{2n}})`$ $`+\mathrm{}+((1)^{ϵ_{2n1}}+(1)^{ϵ_{2n}})].`$ We make now the change of variables: $`\omega _1=ϵ_1+(ϵ_3+ϵ_4)+\mathrm{}+(ϵ_{2n1}+ϵ_{2n})`$ (156) $`\omega _2=ϵ_2+(ϵ_3+ϵ_4)+\mathrm{}+(ϵ_{2n1}+ϵ_{2n})`$ (157) $`\omega _3=ϵ_3+(ϵ_5+ϵ_6)+\mathrm{}+(ϵ_{2n1}+ϵ_{2n})`$ (158) $`\omega _4=ϵ_4+(ϵ_5+ϵ_6)+\mathrm{}+(ϵ_{2n1}+ϵ_{2n})`$ (159) $`\mathrm{}`$ (160) $`w_{2n1}=ϵ_{2n1}`$ (161) $`w_{2n}=ϵ_{2n}.`$ (162) Notice the identity: $$(1)^{_{i=1}^N\omega _i}=(1)^{_{i=1}^Nϵ_i}$$ (163) that we are going to use shortly. With the change of variables (156), instead of the expression (155), we get: $$C_D^{(N)}=\frac{1}{2}\underset{i=1}{\overset{N}{}}(1)^{\omega _i}=\frac{1}{2}\underset{i=1}{\overset{N}{}}\tau _i^z=S_{(N)}^z$$ (164) From equation (164) it results that the spectrum of $`C^{(N)}`$ for $`N=2n`$ is the same as that of the $`z`$-component of the total spin $`S^z`$ for $`2n`$ spins $`\frac{1}{2}`$ . Therefore the eigenvalues are $`nm`$ ($`m=0,1,\mathrm{},N`$) with a degeneracy given the binomial coefficient $`C_N^{Nm}`$. We now consider the case $`q1`$. We are going to use the recurrence relation (119). We first make a change of basis (see equations (121)-(129)) in the four-state chain with n sites: $$(1,1)1,(2,2)2(1,2)3,(2,1)4,$$ (165) and denote by $`u_i^{(k)}`$, the basis vector on the $`k`$-th site having the $`i`$-th ($`i=1,2,3,4`$) component equal to one and the others zero. In this basis $`E`$ and $`F`$ have the expressions: $`E=\left(\begin{array}{cccc}0& 1& 0& 0\\ 1& 0& 0& 0\\ 0& 0& 0& 0\\ 0& 0& 0& 0\end{array}\right),F=\left(\begin{array}{cccc}q& 0& 0& 0\\ 0& q^1& 0& 0\\ 0& 0& 1& 0\\ 0& 0& 0& 1\end{array}\right).`$ (174) In the same basis, for $`N=2`$, one has $$C^{(2)}=E_1$$ (175) and the eigenvalues (eigenfunctions) are: $$1,\left[\frac{u_1^{(1)}+u_2^{(2)}}{\sqrt{2}}\right];1,\left[\frac{u_1^{(1)}u_2^{(2)}}{\sqrt{2}}\right];0,\left[u_3^{(1)}\right];0,\left[u_4^{(1)}\right].$$ (176) Assume that $`\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}`$ written in the basis $`u_{\alpha _1}^{(1)}u_{\alpha _2}^{(2)}\mathrm{}u_{\alpha _n}^{(n)}(\alpha _i=1,2,3,4),`$ (177) is an eigenfunction of $`C^{(N)}`$ corresponding to the eigenvalue $`\mathrm{\Lambda }`$. We now consider the four wave functions $$\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_i^{(n+1)}(i=1,2,3,4),$$ (178) and act with $`C^{(N+2)}`$ on them using the relation (143). We obtain: $`C_\mathrm{\Lambda }^{(N+2)}\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_1^{(n+1)}`$ $`=`$ $`\mathrm{\Lambda }q\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_1^{(n+1)}+\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_2^{(n+1)}`$ (179) $`C_\mathrm{\Lambda }^{(N+2)}\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_2^{(n+1)}`$ $`=`$ $`\mathrm{\Lambda }q^1\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_2^{(n+1)}+\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_1^{(n+1)}`$ (180) $`C_\mathrm{\Lambda }^{(N+2)}\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_3^{(n+1)}`$ $`=`$ $`\mathrm{\Lambda }\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_3^{(n+1)}`$ (181) $`C_\mathrm{\Lambda }^{(N+2)}\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_4^{(n+1)}`$ $`=`$ $`\mathrm{\Lambda }\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}u_4^{(n+1)}.`$ (182) Two of the wave functions (178) for $`i=3`$ and 4 are therefore eigenfunctions of $`C^{(N+2)}`$ corresponding to the same eigenvalue $`\mathrm{\Lambda }`$. One obtains also the two other eigenvalues: $$\mathrm{\Omega }^\pm =\frac{1}{2}[\mathrm{\Lambda }(q+q^1)\pm \sqrt{\mathrm{\Lambda }^2(qq^1)^2+4}].$$ (183) Notice that if $`\mathrm{\Lambda }=[m]_q,`$ then $`\mathrm{\Omega }^\pm =[m\pm 1]_q`$, where we have used the notation (38). The eigenfunctions corresponding to the eigenvalues $`\mathrm{\Omega }^\pm `$ are: $$\mathrm{\Psi }_\mathrm{\Lambda }^{(N)}(u_1^{(n+1)}+(\mathrm{\Omega }^{(\pm )}q\mathrm{\Lambda })u_2^{(n+1)}).$$ (184) Using the eigenvalues and eigenfunctions (176) for $`N=2`$ and the recurrence relations (179)-(183) and (184) one can get all the eigenvalues and eigenfunctions of $`C^{(N)}`$ for any even $`N`$. The eigenvalues are $$[nm]_q(m=0,\mathrm{},N)$$ (185) with a degeneracy $$C_N^{Nm}.$$ (186) For $`q=1`$ one recovers the spectrum given by $`S_{(N)}^z`$ ( see equation (164)). Notice that the similarity transformations used here are even matrices (see Sec.3.2), therefore the supertrace operation define by eq.(64) stays valid. b) N=2n+1 We start again with $`q=1`$ and from the definition of $`C^{(2n+1)}`$ we have: $$C^{(2n+1)}=C^{(2n)}+\frac{1}{2}\underset{k=1}{\overset{2n}{}}\sigma _k^z.$$ (187) Using the equations (152) and (163)-(164), we get the diagonal form of $`C^{(2n+1)}`$: $$C_D^{(2n+1)}=S_{(2n)}^z+\frac{1}{2}(1)^{n+S_{(2n)}^z}$$ (188) which has obviously the spectrum $$\frac{N}{2}2m(m=0,1,2,\mathrm{},n)$$ (189) with a degeneracy: $$C_N^{2m}.$$ (190) We now consider the case $`q1`$ . Instead of (187) one has: $$C^{(2n+1)}=C^{(2n)}+\frac{1}{s+s^1}F_1F_2\mathrm{}F_n.$$ (191) The recurrence relation (143) stays valid as well as (179)-(183) and (184). As opposed to $`N=2n`$ where we use the recurrence relations starting with $`C^{(2)}`$, for $`N=2n+1`$ we start with $`C^{(3)}`$. In the basis $`u_i^{(1)}(i=1,2,3,4)`$, the matrix $`C^{(3)}`$ is $`C^{(3)}={\displaystyle \frac{1}{s+s^1}}\left(\begin{array}{cccc}s^2& (s+s^1)& 0& 0\\ (s+s^1)& s^2& 0& 0\\ 0& 0& 1& 0\\ 0& 0& 0& 1\end{array}\right),`$ (196) having obviously the following eigenvalues (eigenfunctions): $`\left[{\displaystyle \frac{3}{2}}\right]_q,\left[{\displaystyle \frac{1}{\sqrt{1+q}}}(\sqrt{q}u_1^{(1)}+u_2^{(1)})\right];`$ $`\left[{\displaystyle \frac{1}{2}}\right]_q,\left[{\displaystyle \frac{1}{\sqrt{1+q}}}(u_1^{(1)}\sqrt{q}u_2^{(1)})\right];`$ (197) $`\left[{\displaystyle \frac{1}{2}}\right]_q,\left[u_3^{(1)}\right];`$ $`\left[{\displaystyle \frac{1}{2}}\right]_q\left[u_4^{(1)}\right].`$ (198) The spectrum of $`C^{(2n+1)}`$ is therefore: $$\left[\frac{N}{2}2m\right]_q(m=0,1,\mathrm{},n)$$ (199) with a degeneracy: $$C_{2n+1}^{2m}.$$ (200) From the spectra of the matrix $`C`$, which plays the role of a transfer matrix (see equation (30)), one can derive the mass spectra (the inverse of the correlation lengths) using equations (185) and (199). For $`N=2n`$ we have $$M_m=\mathrm{ln}\frac{[\frac{N}{2}]_q}{[\frac{N}{2}m]_q}(m=1,\mathrm{},n2)$$ (201) and for $`N=2n+1`$ we obtain: $$M_m=\mathrm{ln}\frac{[\frac{N}{2}]_q}{[\frac{N}{2}2m]_q}(m=1,\mathrm{},n).$$ (202) Therefore the system is always massive. It is interesting to note that in the large $`N`$ limit, for $`q=1`$, one obtains (see equations (201)-(202)): $$\underset{N\mathrm{}}{lim}\frac{N}{2}M_m=m(N=2n);\underset{N\mathrm{}}{lim}\frac{N}{2}M_m=2m.(N=2n+1).$$ (203) This implies that in the $`N\mathrm{}`$, all correlation lengths diverge. Looking at the expressions (203) and having in mind that in conformal invariant theories one has similar expressions with $`N`$ substituted by $`L`$ ($`N`$ of $`O(N)`$ replacing $`L`$, the size of the system, of the conformal invariant quantum chain) we would expect some similarity between both physics. The analogy, however is not so simple since the degeneracy of the level m also diverges (see equations (186) and (200)). An explicit calculation of the correlation functions in the large $`N`$ limit, which we didn’t do, will clarify the issue. It is interesting to notice that for $`O(3)`$, spin $`S`$, $`(2S+1)`$-state quantum chain, VBS gives for the the smallest mass $`M_1`$, the following large $`S`$ behavior : $$\text{lim}_S\mathrm{}\frac{S}{2}M_1=1$$ (204) Comparing the equations (203) with (204) we learn that in the asymptotic cases, the largest correlation length is given essentially by the number of states of the chain. ## VI Conclusions We have considered $`q`$-deformed $`O(N)`$ symmetric, $`N`$-state quantum chains defined by Hamiltonians given by equation. (1), (35) and (60). The symmetry is unbroken for free boundary conditions. For $`q1`$ the quantum group symmetry is broken for periodic boundary conditions. For $`q=1`$, no symmetry might be left because of boundary terms which can be chosen as described in Sec.4. Using algebraic methods, the ground-state wave functions for these quantum chains are known exactly for periodic, free and non-diagonal boundary conditions, they all correspond to energy zero. The wave functions are obtained using $`q`$-deformed Clifford algebras. These generalizes the construction of Affleck et al . Using the trace and supertrace operation in an auxiliary space, for $`N`$ even and periodic boundary conditions, one obtains two ground-states one for momentum zero and one for momentum $`\pi `$. This implies that even for a finite number of sites and periodic boundary conditions, the ground state is degenerate. For $`N`$ odd one obtains only translationally invariant ground-states. For free boundary conditions the degeneracy of the ground-state is $`2^{N1}`$. This degeneracy is lifted by boundary terms. We have shown how to compute correlation functions and have derived all the correlation lengths. They are finite and diverge only for $`q=1`$ and $`L\mathrm{}`$. What is the physical relevance of our results? For $`N=4`$ we have shown in the Appendix A how the chain can be mapped into the extended Hubbard model . For all values of $`N`$ one can map our quantum chains for obvious reasons into various ladder models writing the on-rung interaction as a two-site interaction. If what one obtains is physically interesting remains to be seen. On the other hand the wave functions we obtain can be used as trial ground-state for more realistic models . Can the procedure described here be extended to other quantum chains? The answer is yes. One can consider $`q`$-deformed $`Sp(N)`$ symmetric chains. In this case instead of the Clifford algebra one gets the $`q`$-deformed Heisenberg algebra as a tool to compute the wave functions. One can go even one step further and take quantum chains with the superalgebra $`Osp(M/N)`$ as symmetry. In this case the algebra one uses to construct the wave functions is a combination of the Clifford and Heisenberg algebras. These extensions are straightforward. Again, it is an open question if these extensions are interesting from a physical point of view. Last but not least, very simple quadratic algebras were discussed above, if more interesting ones (with $`X_\alpha `$ and $`Y_\alpha `$ in equations (6) and (7) unequal to zero) find their use in equilibrium problems, remains to be seen. They do in non-equilibrium problems. ###### Acknowledgements. This work was supported in part by Conselho Nacional de Desenvolvimento Científico e Tecnológico - CNPq - Brazil, Fundação de Amparo à Pesquisa do Estado de São Paulo, and also DAAD - germany. VR wants to thank Alexander Nersesyan and Michele Fabrizio for discussions and also his colleagues at SISSA for hospitality under the TMR grant ERBFMRXCT 960012. ## Appendix A. Fermionic formulation of the $`O(4)`$ quantum chain In this appendix we are going to present explicitly the Hamiltonian that corresponds to the $`N=4`$ case. From (35) and (39) the Hamiltonian is $$H=\underset{k}{}H_k,$$ (A.1) $`H_k=P_k^{(+)}={\displaystyle \underset{\alpha ,\beta ,\gamma ,\delta =1}{\overset{4}{}}}\mathrm{\Gamma }_{\gamma \delta }^{\alpha \beta }E_k^{\gamma \alpha }E_{k+1}^{\delta \beta }=`$ (A.2) $`{\displaystyle \frac{1}{q+q^1}}\{(q+q^1)[E_k^{11}E_{k+1}^{11}+E_k^{22}E_{k+1}^{22}+E_k^{33}E_{k+1}^{33}`$ (A.3) $`+E_k^{44}E_{k+1}^{44}]+q[E_k^{11}E_{k+1}^{22}+E_k^{11}E_{k+1}^{33}+E_k^{22}E_{k+1}^{44}+E_k^{33}E_{k+1}^{44}]+q^1[E_k^{33}E_{k+1}^{11}+E_k^{44}E_{k+1}^{33}`$ (A.4) $`+E_k^{44}E_{k+1}^{22}+E_k^{22}E_{k+1}^{11}]+[E_k^{21}E_{k+1}^{12}+E_k^{12}E_{k+1}^{21}+E_k^{31}E_{k+1}^{13}+E_k^{13}E_{k+1}^{31}+E_k^{42}E_{k+1}^{24}`$ (A.5) $`+E_k^{24}E_{k+1}^{42}+E_k^{43}E_{k+1}^{34}+E_k^{34}E_{k+1}^{43}]+\alpha _3[E_k^{22}E_{k+1}^{33}+E_k^{33}E_{k+1}^{22}+E_k^{14}E_{k+1}^{41}+E_k^{23}E_{k+1}^{32}`$ (A.6) $`+E_k^{32}E_{k+1}^{23}+E_k^{41}E_{k+1}^{14}]+\alpha _1E_k^{11}E_{k+1}^{44}+\alpha _5E_k^{44}E_{k+1}^{11}+\alpha _2[E_k^{31}E_{k+1}^{24}+E_k^{21}E_{k+1}^{34}`$ (A.7) $`+E_k^{12}E_{k+1}^{43}+E_k^{13}E_{k+1}^{42}]+\alpha _4[E_k^{42}E_{k+1}^{13}+E_k^{43}E_{k+1}^{12}+E_k^{34}E_{k+1}^{21}+E_k^{24}E_{k+1}^{31}]\},`$ (A.8) where $$\alpha _1=\frac{q^3}{1+q^2},\alpha _2=\frac{q^2}{1+q^2},\alpha _3=\frac{q}{1+q^2},\alpha _4=\frac{1}{1+q^2},\alpha _5=\frac{q^1}{1+q^2}.$$ (A.9) It is also interesting to rewrite (A.2) in terms of spin-$`\frac{1}{2}`$ creation and annihilation fermion operators on the lattice. This is done by making the following correspondence between the basis $`|\alpha >_j,\alpha =1,2,3,4`$, in (A.2), at each lattice point $`j`$, and the Fock representation $`|1>_j`$ $``$ $`|0>_j=|>_j,|2>_jc_{j,+}^+|0>_j=|>_j,`$ (A.10) $`|3>_j`$ $``$ $`c_{j,}^+|0>_j=|>_j,|4>_jc_{j,+}^+c_{j,}^+|0>_j=|>_j.`$ (A.11) Using this fermionic basis the Hamiltonian density (A.2) is given by $`H_k`$ $`=`$ $`{\displaystyle \frac{1}{q+q^1}}\{{\displaystyle \underset{\sigma =+,}{}}(c_{k,\sigma }^+c_{k+1,\sigma }+\text{h}.c.)(1+t_{\sigma 1}n_{k,\sigma }+t_{\sigma 2}n_{k+1,\sigma }+t_\sigma ^{}n_{k,\sigma }n_{k+1,\sigma })`$ (A.16) $`+J(\stackrel{}{S_k}\stackrel{}{S}_{k+1}n_kn_{k+1}/4)+t_p(c_{k,+}^+c_{k,}^+c_{k+1,+}c_{k+1,}+\text{h}.c.)+(q+q^1)qn_k`$ $`q^1n_{k+1}+U_ln_{k,+}n_{k,}+U_rn_{k+1,+}n_{k+1,}+{\displaystyle \underset{\sigma ,\sigma ^{}=+,}{}}V_{\sigma ,\sigma ^{}}n_{k,\sigma }n_{k+1,\sigma ^{}}`$ $`+[V_3^{(1)}n_{k,+}n_{k+1,}n_{k+1,+}+V_3^{(2)}n_{k,}n_{k+1,}n_{k+1,+}+V_3^{(3)}n_{k,}n_{k,+}n_{k+1,+}`$ $`+V_3^{(4)}n_{k,}n_{k,+}n_{k+1,}]+V_4n_{k,}n_{k,+}n_{k+1,}n_{k+1,+}\},`$ where $`t_1`$ $`=`$ $`t_{+1}={\displaystyle \frac{q^2+2}{q^2+1}},t_2=t_{+2}={\displaystyle \frac{1+2q^2}{1+q^2}},t_{}^{}=t_+^{}=3,J=2t_p={\displaystyle \frac{2q}{1+q^2}},`$ (A.17) $`U_l`$ $`=`$ $`{\displaystyle \frac{q^3}{1+q^2}},U_r={\displaystyle \frac{q^1}{1+q^2}},V_{++}=V_{}=q+q^1,V_+=V_+={\displaystyle \frac{2q}{1+q^2}}`$ (A.18) $`V_3^{(1)}`$ $`=`$ $`V_3^{(2)}=q^1,V_3^{(3)}=V_3^{(4)}=q,V_4=q+q^1.`$ (A.19) In (A.16) appear the density operators $`n_{k,\sigma }=c_{k,\sigma }^+c_{k,\sigma }`$ and $`n_k=n_{k,+}+n_{k,}`$ at the site $`k`$. The magnetic spin-spin interaction (coupling $`J`$) in (A.16) is derived from the relation $$\underset{\sigma \sigma ^{}}{}c_{k,\sigma }^+c_{k+1,\sigma ^{}}^+c_{k,\sigma ^{}}c_{k+1,\sigma }=2(\stackrel{}{S_k}\stackrel{}{S}_{k+1}n_kn_{k+1}/4)+n_{k,+}n_{k+1,}+n_{k,}n_{k+1,+},$$ (A.20) where $`\stackrel{}{S_k}=\frac{1}{2}\stackrel{}{\sigma _k}`$, and $`\stackrel{}{\sigma }=(\sigma ^x,\sigma ^y,\sigma ^z)`$ are the spin-$`\frac{1}{2}`$ Pauli matrices. The Hamiltonian (A.16) belongs to the class of extended Hubbard models considered in the recent literature . Beyond the magnetic interaction (coupling $`J`$) we also have non-diagonal interactions that correspond to single particle correlated hopping (couplings $`t_{\sigma 1},t_{\sigma 2},t_\sigma ^{};\sigma =\pm `$), as well as pair hopping terms (coupling $`t_p`$). The static interactions are given by the diagonal terms. The couplings $`U_l`$ and $`U_r`$ give us the on-site Coulomb interaction, and the interactions $`V_{\sigma ,\sigma ^{}}`$ ($`\sigma ,\sigma ^{}=\pm `$), $`V_3^{(\alpha )}`$, ($`\alpha =1,\mathrm{},4`$) and $`V_4`$ give us the two- three- and four-body static interactions, respectively. We should notice that the Hamiltonian (A.16) conserves separately the total number of up spins $`n_+`$ and down spins $`n_{}`$. Consequently for free boundary consitions we may construct, using the the algebraic method, zero-energy eigenfunctions $`\mathrm{\Psi }_{n_+,n_{}}`$, for each sector labelled by $`n_+`$ and $`n_{}`$ ($`n_+,n_{}=0,1,\mathrm{},L`$), i. e., $$\mathrm{\Psi }_{n_+,n_{}}=𝒫_{n_+,n_{}}\left[\underset{_{k=1}}{\overset{L}{}}\left(x_1+x_2c_{k,+}^++x_3c_{k,}^++x_4c_{k,+}^+c_{k,}^+\right)|0>_k\right],$$ (A.21) where $`𝒫_{n_+,n_{}}`$ projects out states which do not have $`n_+`$ spins $`\sigma =+`$ and $`n_{}`$ spins $`\sigma =`$ (see equation (70)). ## Appendix B. Correlation functions for parity violating operators ($`N`$ even) We would like to show how to compute the correlation functions $$\xi _{r,s}=\frac{<0,|P_rQ_s|0,+>}{Z_{,+}}$$ (B.1) where $$Z_{,+}=<0,|0,+>$$ (B.2) which appear for $`N`$ and $`L`$ even and periodic boundary conditions when the vacuum is degenerate. Here $$|0,+>=\text{Tr}(x_{\alpha _1}\mathrm{}x_{\alpha _L})u_{\alpha _1}\mathrm{}u_{\alpha _L}$$ (B.3) corresponds to the parity $`+`$, momentum zero wave function, $$<0,|=\text{Str}(y_{\beta _1}\mathrm{}y_{\beta _L})u_{\beta _1}^T\mathrm{}u_{\beta _L}^T=\text{Tr}(Jy_{\beta _1}\mathrm{}y_{\beta _L})u_{\beta _1}^T\mathrm{}u_{\beta _L}^T$$ (B.4) corresponds to the parity $``$, momentum $`\pi `$ wave function. The matrix $`J`$ is defined in equation (61). The action of the operators $`P`$ and $`Q`$ is shown equation (25). For obvious reasons, one of the two operators $`P`$ or $`Q`$ has to break parity. It is easy to show, using the definitions given by equations (25) and (27) that we have $$<0,|P_rQ_s|0,+>=\text{Tr}(DC^{r2}PC^{sr1}QC^{Ls})$$ (B.5) and $$<0,|0,+>=\text{Tr}(DC^{L1})$$ (B.6) where $$D=\underset{\alpha =1}{\overset{N}{}}x_\alpha Jy_\alpha .$$ (B.7) Obviously the correlation lengths appearing for this type of correlation functions are the same as for the parity conserving operators where one computes quantities like $`<0,+|\mathrm{}|0,+>`$ or $`<0,|\mathrm{}|0,>`$.
warning/0004/hep-ph0004085.html
ar5iv
text
# 1 Introduction ## 1 Introduction Proposed neutrino factories offer unique possibilities to improve the knowledge about neutrino masses, leptonic mixings and CP violation. The leptonic sector is not plagued by hadronic uncertainties such that vacuum neutrino oscillation allows to determine many parameters quite precisely. Vacuum neutrino oscillation is however only sensitive to mass squared differences thus the sign of $`\mathrm{\Delta }m_{31}^2`$ can not be determined allowing therefore at the moment different scenarios for the ordering of mass eigenvalues. It has however been pointed out recently that the sign of $`\mathrm{\Delta }m_{31}^2`$ can be determined from the $`\nu _e\nu _\mu `$ and $`\overline{\nu }_e\overline{\nu }_\mu `$ appearance rates via matter effects in very long baseline neutrino oscillation experiments at neutrino factories. We discuss and quantify in this paper further possibilities to determine the sign of $`\mathrm{\Delta }m_{31}^2`$ via matter effects and $`\mathrm{sin}^22\theta _{13}`$ by considering the appearance channels and/or the $`\nu _\mu \nu _\mu `$ and $`\overline{\nu }_\mu \overline{\nu }_\mu `$ disappearance channels in different scenarios with and without muon charge identification. Altogether there are four possibilities. First, without charge identification one can not study wrong sign muon events (i.e. the appearance channels), but one can operate a neutrino factory already both with $`\mu ^{}`$ and $`\mu ^+`$ beams and analyze the differences in the combined muon neutrino and antineutrino event rate spectrum. This is useful if charge identification is not available or not operative in an initial stage of the experiment. We show that this leads already to a quite good sensitivity since the effects in the combined appearance and disappearance channels are for a baseline of $`7332`$ km about the same size and go in the same direction. Next, with muon charge identification, one can study the $`\nu _\mu \nu _\mu `$ disappearance and the $`\nu _e\nu _\mu `$ appearance channels separately. For a baseline between $`2800`$ km and $`7332`$ km we find that the significance of these result is essentially unchanged compared to 7332 km. The point is that the total event rates decrease slower than the vacuum rates leading for a large range to a statistically constant result. We use therefore in this paper mostly a baseline of $`7332`$ km since for this larger baseline there is further information in the disappearance channels. A third option would be to use the disappearance channel alone, this provides however not much extra information. Last, the best method is achieved with muon charge identification by combining all available information of the $`\nu _\mu \nu _\mu `$ disappearance and $`\nu _e\nu _\mu `$ appearance channels. We show for our baseline of $`7332`$ km the sensitivity which can be achieved. The basic effect which allows to extract the sign of $`\mathrm{\Delta }m_{31}^2`$ comes from coherent forward scattering of electron neutrinos in matter which leads to effective masses and mixings different from vacuum. We will discuss in this paper a full three neutrino framework in the limit where the small solar mass splitting can be neglected in vacuum, i.e. $`\mathrm{\Delta }m_{21}^2=0`$. This degeneracy is however destroyed in matter and we need therefore a full three neutrino description. The basic mechanism which allows the extraction of the sign of $`\mathrm{\Delta }m^2`$ via matter effects can, however, already be seen in a simplified 2x2 picture with two mass eigenvalues $`m_i`$, $`m_j`$, $`\mathrm{\Delta }m^2=m_j^2m_i^2`$ and one mixing angle $`\theta `$ only. One obtains then the well known relations for the parameters in matter $$\mathrm{\Delta }m_m^2=\mathrm{\Delta }m^2C_\pm =\mathrm{\Delta }m^2{}_{}{}^{^+}\sqrt{\left(\frac{A}{\mathrm{\Delta }m^2}\mathrm{cos}2\theta \right)^2\pm \mathrm{sin}^22\theta };\mathrm{sin}^22\theta _m=\mathrm{sin}^22\theta C_\pm ^2,$$ (1) where the neutrino energy $`E`$ enters via the matter term $$A2EV=\pm \frac{2\sqrt{2}G_FY\rho E}{m_n}.$$ (2) The sign of the matter term $`A`$ is for electron neutrinos (antineutrinos) traveling inside the earth positive (negative) and leads thus to different corrections for neutrinos ($`C_+`$) and antineutrinos ($`C_{}`$). Matter effects modify the 2x2 vacuum transition probabilities $$P(\nu _i\nu _j)=\mathrm{sin}^22\theta \mathrm{sin}^2\left(\frac{\mathrm{\Delta }m^2L}{E}\right)\mathrm{and}P(\nu _{i,j}\nu _{i,j})=1P(\nu _i\nu _j),$$ (3) since in matter one must replace $`\mathrm{sin}^22\theta \mathrm{sin}^22\theta _m`$ and $`\mathrm{\Delta }m^2\mathrm{\Delta }m_m^2`$. The matter corrections in eq. (1) depend for a given neutrino species (i.e. given A) clearly on the sign of $`\mathrm{\Delta }m^2`$ allowing thus in principle an extraction of this sign. One could, for example, compare neutrinos with antineutrinos, since the matter corrections go then in opposite direction inducing an asymmetry. The biggest matter effects (and therefore the best sensitivity to the sign of $`\mathrm{\Delta }m^2`$) occur when eq. (1) becomes “resonant” for $$A=\mathrm{\Delta }m^2\mathrm{cos}2\theta .$$ (4) This condition can be fulfilled for a given sign of $`\mathrm{\Delta }m^2`$ either for neutrinos or for antineutrinos at a specific resonance energy, but optimization will be more complicated in a real experiment with an energy spectrum and other free parameters. Nevertheless eq. (4) leads already to a rough approximation for a “optimal mean neutrino energy”. In earth eq. (4) gives for small $`\theta `$, i.e. for $`\mathrm{cos}2\theta =1`$ roughly the relation $$E_{opt}=15\mathrm{GeV}\left(\frac{\mathrm{\Delta }m_{31}^2}{3.5\times 10^3\mathrm{eV}^2}\right)\left(\frac{2.8\mathrm{g}/\mathrm{cm}^3}{\rho }\right),$$ (5) i.e. an optimal energy of about 15 GeV in the crust of the earth with an average effective density $`\rho 2.8\mathrm{g}/\mathrm{cm}^3`$ going down to about 10 GeV for paths crossing the earth at a distance of 7332 km with $`\rho 4.2\mathrm{g}/\mathrm{cm}^3`$. Our paper is organized as follows. First we develop in section 2 analytic expressions for neutrino oscillations in matter in a suitable approximation. In section 3 we describe the experimental framework on which our numerical results are based. Section 4 describes the effects of matter on the total rates and the used statistical methods are outlined in chapter 5. Section 6 contains our results on the sensitivity to the magnitude and sign of $`\mathrm{\Delta }m_{31}^2`$ and for $`\mathrm{sin}^22\theta _{13}`$. ## 2 Analytic Description of three Neutrino Mixing in Matter The mixing of three neutrinos can be described via $$|\nu _l>=\underset{k=1}{\overset{3}{}}U_{lk}|\nu _k>,l=e,\mu ,\tau ,$$ (6) where $`|\nu _l>`$ corresponds to the neutrino flavour state $`\nu _l`$ and $`|\nu _k>`$ corresponds to the neutrino mass eigenstate $`\nu _k`$ with eigenvalues $`m_k`$, $`m_km_j`$, $`kj=1,2,3`$. $`U`$ is a $`3\times 3`$ unitary leptonic mixing matrix which we parameterize as $$\left(\begin{array}{ccc}U_{e1}& U_{e2}& U_{e3}\\ U_{\mu 1}& U_{\mu 2}& U_{\mu 3}\\ U_{\tau 1}& U_{\tau 2}& U_{\tau 3}\end{array}\right)=\left(\begin{array}{ccc}c_{12}c_{13}& s_{12}c_{13}& s_{13}e^{i\delta }\\ s_{12}c_{23}c_{12}s_{23}s_{13}e^{i\delta }& c_{12}c_{23}s_{12}s_{23}s_{13}e^{i\delta }& s_{23}c_{13}\\ s_{12}s_{23}c_{12}c_{23}s_{13}e^{i\delta }& c_{12}s_{23}s_{12}c_{23}s_{13}e^{i\delta }& c_{23}c_{13}\end{array}\right)$$ (7) where $`c_{ij}=\mathrm{cos}\theta _{ij}`$ and $`s_{ij}=\mathrm{sin}\theta _{ij}`$. The mixing matrix contains in general further CP-violation phases which do not enter in neutrino oscillation and are therefore not specified . The angles are assumed to be in the ranges $`0\theta _{12},\theta _{23},\theta _{13}<\pi /2,0\delta <2\pi `$. The squared mass splittings $`\mathrm{\Delta }m^2=m_j^2m_i^2`$ show the hierarchy $`\mathrm{\Delta }m_{21}^2|\mathrm{\Delta }m_{31}^2|`$ as discussed in more detail below and allow for $`\mathrm{\Delta }m_{21}^210^4\mathrm{eV}^2`$, $`E1\mathrm{GeV}`$ and $`L10^4`$ km the approximation where $`\mathrm{\Delta }m_{21}^2=0`$ in vacuum. We have thus two almost degenerate mass eigenvalues and for $`\mathrm{\Delta }m_{31}^2>0`$ ($`\mathrm{\Delta }m_{31}^2<0`$) these almost degenerate eigenvalues $`m_1`$ and $`m_2`$ are lighter (heavier) than $`m_3`$ (see fig. 1). The usual oscillation formulas depend only on $`\mathrm{\Delta }m_{31}^2`$ and it is thus impossible to discriminate between the two schemes. Matter effects and CP-violating effects can lift this degeneracy in future measurements in very long baseline experiments. We can assume therefore for the following discussion that there is only one non-vanishing quadratic mass splitting $`\mathrm{\Delta }m_{32}^2=\mathrm{\Delta }m_{31}^2=\mathrm{\Delta }m^2`$ in vacuum. To understand this case in matter we can write the mixing matrix eq. (7) as a sequence of rotations $`U=R(\theta _{23})R(\theta _{13})R(\theta _{12})`$ where the CP-phase $`\delta `$ is included in eq. (7) by a complex rotation in 13-subspace, i.e. by replacing $`s_{13}s_{13}e^{i\delta }`$. Alternatively the CP-phase could be included by making any of the three rotations complex. This becomes useful when we write the full Hamiltonian in matter with the help of $`U`$ in flavour basis<sup>1</sup><sup>1</sup>1A constant common neutrino mass can be separated leading to an overall phase. $$=\frac{1}{2E}R(\theta _{23})\left[R(\theta _{13})\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& 0\\ 0& 0& \mathrm{\Delta }m^2\end{array}\right)R(\theta _{13})^T+\left(\begin{array}{ccc}A& 0& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right)\right]R(\theta _{23})^T.$$ (8) Here use has been made of the fact $`R(\theta _{12})`$, which operates in 12-subspace, drops out in the limit where $`\mathrm{\Delta }m_{21}^2=0`$. Since we could use a convention where the CP-phase would be defined by a complex rotation in 12-subspace we can immediately see that the CP-phase drops out as well. Furthermore $`R(\theta _{23})`$ operates in the 23-subspace and commutes therefore with the matter term. We see that matter effects are confined in eq. (8) inside the square brackets, which implies that matter affects in the $`\mathrm{\Delta }m_{21}^2=0`$ limit only the 13-subspace. We can readily use the standard two neutrino parameter mapping given in eqs. (1) for the 13-subspace and obtain in matter $`\mathrm{\Delta }m_{31,m}^2=\mathrm{\Delta }m^2C_\pm `$ and $`\mathrm{sin}^22\theta _m=\mathrm{sin}^22\theta /C_\pm ^2`$, where $`\mathrm{\Delta }m^2\mathrm{\Delta }m_{31}^2`$ enters into $`C_+`$ defined in eq. 1. Note, however, that $`\mathrm{\Delta }m_{32}^2`$ and $`\mathrm{\Delta }m_{21}^2`$ become in matter in this way also energy dependent, namely $$\mathrm{\Delta }m_{32,m}^2=\mathrm{\Delta }m^2C_\pm ^{}=\frac{\mathrm{\Delta }m^2(C_\pm +1)+A}{2};\mathrm{\Delta }m_{21,m}^2=\mathrm{\Delta }m^2C_\pm ^{\prime \prime }=\frac{\mathrm{\Delta }m^2(C_\pm 1)A}{2}.$$ (9) An important consequence is that the mass degeneracy between the first two eigenvalues is destroyed in matter. Thus even if we can approximate $`\mathrm{\Delta }m_{21}^2=0`$ in vacuum we need for oscillation in matter the full three neutrino oscillation formulae. The only simplification left is that one can set in the mixing matrix $`\theta _{12}=0`$ and $`\delta =0`$. These matter induced parameter mappings must now be inserted into the oscillation formulae for three neutrinos. Defining as shorthands $`D_{ab}=e^{iE_at}\delta _{ab}`$, $`J_{ij}^{lm}:=U_{li}U_{lj}^{}U_{mi}^{}U_{mj}`$ and $`\mathrm{\Delta }_{ij}:=\mathrm{\Delta }m_{ij}^2L/4E`$ the transition probabilities $`P(\nu _l\nu _m)`$ from flavour $`l`$ to flavour $`m`$ in vacuum can be written as $$P(\nu _l\nu _m)=\left|l|UDU^+|m\right|^2=\delta _{lm}4\underset{i>j}{}\mathrm{Re}J_{ij}^{lm}\mathrm{sin}^2\mathrm{\Delta }_{ij}2\underset{i>j}{}\mathrm{Im}J_{ij}^{lm}\mathrm{sin}2\mathrm{\Delta }_{ij}.$$ (10) As described above we can set in matter $`\theta _{12}=0`$ and $`\delta =0`$ resulting in $`P(\nu _\mu \nu _e)=`$ $`\mathrm{sin}^2\theta _{23}\mathrm{sin}^22\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{31}),`$ (11a) $`P(\nu _\mu \nu _\mu )=`$ $`1\mathrm{sin}^22\theta _{23}\mathrm{sin}^2\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{21})\mathrm{sin}^4\theta _{23}\mathrm{sin}^22\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{31})`$ $`\mathrm{sin}^22\theta _{23}\mathrm{cos}^2\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{32}),`$ (11b) $`P(\nu _\mu \nu _\tau )=`$ $`\mathrm{sin}^22\theta _{23}[\mathrm{sin}^2\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{21}){\displaystyle \frac{1}{4}}\mathrm{sin}^22\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{31})`$ $`+\mathrm{cos}^2\theta _{13}\mathrm{sin}^2(\mathrm{\Delta }_{32})].`$ (11c) To describe oscillation in matter we must insert into these formulae the parameter mappings discussed above, namely the shifted mass eigenvalues ($`\mathrm{\Delta }m_{31,m}^2=\mathrm{\Delta }m^2C_\pm `$ and eqs. (9)) and the shifted 13-mixing angle, $`\mathrm{sin}^22\theta _{13,m}=\mathrm{sin}^22\theta _{13}/C_\pm ^2`$. Note that the relations for antineutrinos are only formally identical due to the opposite sign of the matter term $`A`$ in eq. (1). In vacuum the relations (10a)–(10c) could be further simplified since the tight CHOOZ results can be taken into account. For typically allowed $`\mathrm{\Delta }m_{31}^2`$ values one has $`\mathrm{sin}^22\theta _{13}<0.1`$, while this upper limit is less stringent for $`1.0\times 10^3\mathrm{eV}^2|\mathrm{\Delta }m_{31}^2|<3.0\times 10^3\mathrm{eV}^2`$ where values of $`\mathrm{sin}^22\theta _{13}0.2`$ are allowed. Note however that the effective $`\mathrm{sin}^22\theta _{13}`$ in matter can be much bigger when the resonance condition is fulfilled such that further simplifications of eqs. (10a)–(10c) are potentially dangerous. ## 3 Experimental Framework We use for our study a standard earth matter density profile and we will treat neutrino oscillations in earth by a constant density approximation along each neutrino path. The varying matter profile of the earth implies that the average density is still path dependent. It has been shown recently that this approximation works very well, while a constant density for all paths ignores even the mean density variations in the earth and has sizable errors<sup>2</sup><sup>2</sup>2Note also that care should be taken about the core of the earth. Either one should avoid paths passing the core or one should take core effects into account. Specific core effects become however small when the oscillation wavelength becomes bigger than the size of the core.. We avoid the core of the earth and take in our study distances up to 7332 km. We assume furthermore a neutrino factory which can operate with $`\mu ^{}`$ or $`\mu ^+`$ beams leading to $`\nu _\mu +\overline{\nu }_e`$ or $`\overline{\nu }_\mu +\nu _e`$ neutrino beams resulting from $`N_{\mu ^\pm }`$ muon decays. The detector with $`N_{\text{kT}}`$ kilotons is assumed to be able to measure muons above an assumed threshold of $`5`$ GeV with an efficiency $`ϵ_\mu `$. The muons may come from the dominant $`\nu _\mu \nu _\mu `$ disappearance channel or from the subdominant $`\nu _e\nu _\mu `$ appearance channel with or without charge identification. Muons coming from the subdominant $`\nu _\mu \nu _\tau `$ or $`\nu _e\nu _\tau `$ channels with the $`\tau `$ decaying subsequently to a muon should not be a problem, since they can be separated by kinematical means without loosing too much muon detection efficiency . The inclusion of various other potential backgrounds and detector properties will have some influence on the analysis, but with current understanding this would only result in moderate corrections. More detailed simulations of backgrounds are still in progress . Most limitations due to backgrounds and statistics can however be overcome by increasing the number of muon decays and/or the size of the detector. We do not include a detailed simulation of background effects at the moment. For three neutrinos there is only room for two independent quadratic mass splittings while there are three different results for oscillation. Among these the LSND evidence for oscillation is most controversial and we omit this result therefore in our analysis, while we take the atmospheric result (mostly from SuperKamiokande ) and the solar neutrino deficit (mostly from GALLEX ) as evidence for oscillation. These results can be studied in a simplified 2x2 oscillation picture where the mixing matrix contains only one mixing angle $`\theta `$ with the two flavour transition probabilities in vacuum given in eq. (3). The dominant atmospheric $`\nu _\mu \nu _\tau `$ oscillations imply in this picture a mass splitting $$10^3\mathrm{eV}^2<|\mathrm{\Delta }m_{31}^2|<8.0\times 10^3\mathrm{eV}^2,$$ (12) while the solar neutrino deficit leads to different solutions for $`\mathrm{\Delta }m_{21}^2`$, all with $`|\mathrm{\Delta }m_{21}^2||\mathrm{\Delta }m_{31}^2|`$. These are the solar vacuum oscillation (VO) solution and the large and small mixing angle (LMA and SMA) solar MSW solutions. The above assignment implies thus that $`\mathrm{\Delta }m_{31}^2`$ dominates the atmospheric $`\nu _\mu \nu _\tau `$ oscillation, while $`\mathrm{\Delta }m_{21}^2`$ is most relevant for the VO, SMA MSW and LMA MSW solution of the solar neutrino problem. We will see that our study depends only on $`\mathrm{\Delta }m_{31}^2`$ such that in a very good approximation $`\mathrm{\Delta }m_{21}^2`$ can be set to zero. We use therefore as default initial parameters of our study $$|\mathrm{\Delta }m_{31}^2|=3.510^3;|\mathrm{\Delta }m_{21}^2|=0;\mathrm{sin}^22\theta _{23}=1.0$$ (13) and we will discuss variations of these parameters within the allowed ranges. The event rates of our results depend only on the combination $`N_{\mu ^\pm }N_{\text{kT}}ϵ_\mu `$ and we take for both polarities $`N_{\mu ^\pm }N_{\text{kT}}ϵ_\mu =210^{21}`$. This corresponds for example to a setup with $`N_{\mu ^+}=210^{20}\mu ^+`$ decays, $`N_\mu ^{}=210^{20}\mu ^{}`$ decays, an $`N_{\text{kT}}=10`$ kt iron detector and a muon detection efficiency $`ϵ_\mu =1`$. Unless otherwise mentioned we use a baseline of $`L=7332`$ km. Our analysis is performed at the level of total and differential event rates. We include therefore the charged current neutrino cross sections per nucleon in the detector and the normalized $`\nu _\mu +\overline{\nu }_e`$ and $`\overline{\nu }_\mu +\nu _e`$ beam spectra $`g_{\nu _i}`$ of the neutrino factory $`\sigma _{\nu _\mu }(E)`$ $`=0.6710^{38}E\text{cm}^2/\text{GeV},`$ $`\sigma _{\overline{\nu }_\mu }(E)`$ $`=0.3410^{38}E\text{cm}^2/\text{GeV},`$ (14a) $`g_{\nu _e}(x)`$ $`=g_{\overline{\nu }_e}(x)=12x^2(1x),`$ $`g_{\nu _\mu }(x)`$ $`=g_{\overline{\nu }_\mu }(x)=2x^2(32x),`$ (14b) where $`x=E/E_\mu `$. For a given number of useful muon decays $`N_{\mu ^\pm }`$, detector size $`N_{\text{kT}}`$, efficiency $`ϵ_\mu `$ one obtains for the channel ”i” (where i stands for a proper index uniquely given by the polarity of the muon beam and the considered oscillation channel) the contributions to the differential event rates $$\frac{dn_i}{dE}=\underset{normalization}{\underset{}{\left[N_{\mu ^i}N_{\text{kT}}ϵ_\mu \frac{10^9N_A}{m_\mu ^2\pi }\right]}}\underset{flux}{\underset{}{\left[\frac{E_\mu }{L^2}g_i(E/E_\mu )\sigma _i(E)\right]}}\underset{oscillation}{\underset{}{\left[P_i(E)\frac{}{}\right]}}$$ (15) where $`10^9N_A`$ is the number of nucleons per kiloton in the detector and $`P_i`$ stands for the relevant oscillation probability in matter as discussed above. In cases where different channels contribute the total differential rates are given by the sum of all individual terms. Total rates are obtained by integrating these differential rates from the threshold at $`5`$ GeV to the maximal possible neutrino energy $`E_\mu `$. To understand the event rates we look at the three terms in square brackets on the rhs. of eqs. (15). The first term contains overall factors which are constant in energy, which is only important for the proper event rate normalization. The second term is simply the product of flux times detection cross section which has to be folded with the third term, the oscillation probability in matter. The second term grows initially like $`E^3`$ and reaches a maximum before it goes to zero at $`E=E_\mu `$. An important feature of the second term is that the low energy part is not changed when the muon energy is increased. This implies in a good approximation that the increase in the total event rates by increasing $`E_\mu `$ comes essentially from the high energy tail of the spectrum only. This has also the important consequence that experimental results for a lower muon beam energy $`E_\mu ^{}<E_\mu `$ can in principle be obtained to a good approximation at higher energies by simply removing events with energies above $`E_\mu ^{}`$. ## 4 Total Rates The analytic description above allows now an understanding of the event rates which we calculated numerically and which are presented below in fig. 2. To understand first the total disappearance rates without matter one must look at the folding of the vacuum oscillation probabilities with the flux factor in eq. (15). The oscillatory terms in the oscillation probabilities eqs. (10), namely the $`\mathrm{sin}^2\mathrm{\Delta }_{ij}`$ factors, depend then for fixed $`\mathrm{\Delta }m_{31}^2=\mathrm{\Delta }m^2`$ only on $`E/L`$. The oscillatory behavior vanishes for $`E/L\mathrm{}`$ and the first maximum of the vacuum oscillation occurs for given $`L`$ at an energy $`E_1=2\mathrm{\Delta }m^2L/\pi `$. This means that there will essentially be no effects of oscillation in the total disappearance rates when $`E_1`$ is much smaller than the neutrino energy $`E`$. Since the event rates are dominated by the maximum of the flux factor close to $`E_\mu `$ we can set $`EE_\mu `$ and find no oscillation effects when $`E_\mu `$ is much larger than $`E_1`$. The overall rates will however still decrease with the muon energy $`E_\mu `$ according to the cross sections and fluxes approximately like $`E_\mu ^2`$ as long as $`E_\mu `$ is sufficiently larger than the muon threshold energy of $`5`$ GeV. The first effects of vacuum oscillation show up for lower beam energies when $`EE_\mu `$ is decreased such that $`E_\mu =𝒪(E_1)`$. The oscillation becomes maximal for $`E_\mu E_1`$ while it vanishes again for $`E_\mu E_1/2`$. The effects of oscillation can be seen in the total rates as a dip in the overall $`E_\mu ^2`$ scaling. For lower beam energies there are in principle further dips due to all other oscillation maxima, but this is on our case below the threshold energy of $`5`$ GeV. In order to understand how matter effects influence the disappearance rates one must compare the matter “resonance energy” $`E_{opt}`$ in eq. (5) at roughly $`1015`$ GeV with $`E_1`$ and $`E_\mu E`$. At neutrino energies which are not close to this resonance energy the mixing $`\mathrm{sin}^22\theta _{13,m}`$ approaches quickly its value in vacuum, which is small. One can see immediately from the Hamiltonian in eq. (8) that the dominant $`\nu _\mu \nu _\tau `$ oscillation decouples in the limit where $`\mathrm{sin}^22\theta _{13}=0`$ from the matter effects coming from the first generation such that it is clear that matter effects should be rather localized around the resonance region<sup>3</sup><sup>3</sup>3Note that the mass splittings $`\mathrm{\Delta }m_{ij}^2`$ behave non-trivial in the high energy limit such that the asymptotic properties of eq. (10b) have to be evaluated carefully to reach the same conclusion.. The total appearance rates are given for $`E_\mu E_1`$ by the asymptotic behavior of eq. (10a) weighted with the flux. The increase in the rates with $`E_\mu `$ comes again predominantly from the increased flux at energies close to $`E_\mu `$ such that we need to look again in a good approximation at the asymptotic behaviour of $`E_\mu ^2P(\nu _\mu \nu _e)`$. Expanding $`E_\mu ^2P(\nu _\mu \nu _e)`$ in powers of $`1/E`$, and using the fact that $`VL`$ is numerically small, one finds for the asymptotic appearance rates for small $`\mathrm{sin}^22\theta _{13}`$ $$n\mathrm{sin}^2\theta _{23}\mathrm{sin}^22\theta _{13}(\mathrm{\Delta }m^2)^2L^2\left(1\frac{(\mathrm{\Delta }m^2)^2L^4}{3E_\mu ^2}\right)+\frac{4(\mathrm{\Delta }m^2)^3L^4V}{3E_\mu }.$$ (16) The result shows that the rates approach for $`E_\mu E_1`$ in vacuum a constant which depends only on even powers of $`\mathrm{\Delta }m^2`$. The leading corrections to this constant fall like $`1/E_\mu ^2`$. Turning on matter effects leads in eq. (16) to the third term which falls only like $`1/E_\mu `$ and which depends on $`(\mathrm{\Delta }m^2)^3`$, i.e. on the sign of $`\mathrm{\Delta }m_{31}^2`$. This induces in our case a matter dependent splitting in the appearance rates which is less localized in energy than the matter effects in the disappearance channels. There are thus significant effects even for beam energies which are somewhat away from the resonance energy. Note, however, that the discussion of effects is much more complicated for muon energies of the order or smaller than $`E_1`$. This understanding of matter effects can now be compared with the results of our full numerical calculations. The following figures for the total rates in the different oscillation channels assume $`N_{\mu ^\pm }N_{\text{kT}}ϵ_\mu =210^{21}`$, a fixed baseline of 7332 km, $`|\mathrm{\Delta }m_{31}^2|=3.510^3\mathrm{eV}^2`$, $`\mathrm{sin}^22\theta _{23}=1`$ and different values of $`\mathrm{sin}^22\theta _{13}`$. If the muon beam energy is fixed then we use always $`E_\mu =20\mathrm{GeV}`$. Fig. 2 shows the $`\nu _e\nu _\mu `$ and $`\overline{\nu }_e\overline{\nu }_\mu `$ total appearance rates (dotted lines) as well as the $`\nu _\mu \nu _\mu `$ and $`\overline{\nu }_\mu \overline{\nu }_\mu `$ disappearance rates (solid lines) on a logarithmic scale for maximal $`\mathrm{sin}^22\theta _{13}=0.1`$. These rates are in agreement with the results obtained recently by Barger et. al . The lines appear in pairs of neutrino ($``$) and antineutrino ($``$) channels which differ roughly by a factor two coming from the cross sections eqs. (3). Fig. 2 shows the cases with positive $`\mathrm{\Delta }m_{31}^2`$ ($``$), negative $`\mathrm{\Delta }m_{31}^2`$ ($``$) and no oscillation ($``$). The asymmetry between $``$ and $``$ in the appearance rates, which has been used to extract the sign of $`\mathrm{\Delta }m_{31}^2`$ at smaller baseline , is clearly visible. Note that there are also comparable matter effects in the disappearance channels at these large baselines, which appear only to be smaller due to the logarithmic scale. Fig. 2 shows clearly the important feature that the matter effects in the combined appearance and disappearance muon rates go for a neutrino factory with either $`\nu _\mu +\overline{\nu }_e`$ or $`\overline{\nu }_\mu +\nu _e`$ beams in the same direction. This opens an interesting possibility if muon charge identification is not available to separate the appearance channels (wrong sign muon events) from the disappearance channels (right sign muon events). It is then still possible to measure all muons, i.e. muons with both charges, or in other words the combination of the two oscillation channels, where matter effects add up. The resulting total muon rates of the combined channels are shown for both $`\mu ^{}`$ and $`\mu ^+`$ beams in fig. 3 on a linear scale for $`\mu ^{}`$ and $`\mu ^+`$ muon beams. The figure shows the sizable matter induced splittings in the total muon-neutrino rates ($`\nu _\mu +\overline{\nu }_\mu `$) for the mixing angles $`\mathrm{sin}^22\theta _{13}=(0.1,0.04,0.01)`$. The difference between the two signs of $`\mathrm{\Delta }m_{31}^2`$ influences the total rates up to a factor two or more. The effect comes partly from the appearance rates and partly from the disappearance rates, but always from the resonant channel. This is for positive $`\mathrm{\Delta }m_{31}^2`$ the neutrino component of the beam and for negative $`\mathrm{\Delta }m_{31}^2`$ the antineutrino component. For $`\mu ^+`$ beams the interplay of the two channels is as follows<sup>4</sup><sup>4</sup>4For $`\mu ^{}`$ beams holds the same for a reversed sign of $`\mathrm{\Delta }m_{31}^2`$.: For positive $`\mathrm{\Delta }m_{31}^2`$ the $`\nu _e\nu _\mu `$ appearance channel is resonant leading to an enhancement of the combined rate, while matter corrections are moderate for the $`\overline{\nu }_\mu \overline{\nu }_\mu `$ disappearance channel. For negative $`\mathrm{\Delta }m_{31}^2`$, on the other side, the disappearance channel $`\overline{\nu }_\mu \overline{\nu }_\mu `$ shows a resonant transition of $`\nu _\mu `$ to $`\nu _e`$ resulting in a suppression of the total $`\nu _\mu +\overline{\nu }_\mu `$ rate, while the $`\nu _e\nu _\mu `$ appearance rates are mostly unchanged. The combination of appearance and disappearance channels amplifies thus always the signal. This shows that an analysis of the combined appearance and disappearance channels is interesting and it may be especially important if the separation of right and wrong sign muon events is not available. Figure 4 shows the corresponding differential event rate spectrum of the $`\nu _\mu +\overline{\nu }_\mu `$ induced muon rate over twenty energy bins for a mixing angle $`\mathrm{sin}^22\theta _{13}=0.1`$ (CHOOZ bound). We find that the different matter effects have no specific spectral features which could be used as a basis for cuts in order to amplify the signal relative to the oscillation signal. We will fit therefore in the following differential event rate spectra of data simulated with certain initial parameters and with statistical noise added. We will show how well these parameters can be re-extracted in an appropriate statistical way. We provide therefore in the next section information about the used statistical procedures and we will study then in more detail the statistical significance of the matter effects in the appearance channels, disappearance channels and combinations thereof. We focus on the potential to extract $`\theta _{13}`$, the sign of $`\mathrm{\Delta }m_{31}^2`$ and to test matter effects. ## 5 Statistical Methods and Data Evaluation We numerically simulate data for a given parameter set ($`E_\mu `$, $`\mathrm{\Delta }m_{31}^2`$, $`\mathrm{sin}^22\theta _{23}`$, $`\mathrm{sin}^22\theta _{13}`$) and add Poisson distributed fluctuations to the energy bins. The analysis of these “data” is done according to the procedure proposed by the Particle Data Book for Poisson distributed quantities. Confidence levels (CL) are calculated by using $$\chi ^2=\underset{i=1}{\overset{n}{}}\left(2\left[n_i^{\mathrm{th}}n_i^{\mathrm{obs}}\right]+2n_i^{\mathrm{obs}}\mathrm{log}\frac{n_i^{\mathrm{obs}}}{n_i^{\mathrm{th}}}\right).$$ (17) We assume symmetric operation of the neutrino factory in both polarities, i.e. $`\mu ^+`$ and $`\mu ^{}`$ beams together with a detector such that $`N_{\mu ^+}N_{\text{kT}}ϵ_\mu =N_\mu ^{}N_{\text{kT}}ϵ_\mu =210^{21}`$. The total $`\chi ^2`$ is given by $$\chi ^2=\chi _{\mu ^+}^2+\chi _\mu ^{}^2.$$ (18) If there is no charge identification then $`n_i=(n_{\mu ^+})_i+(n_\mu ^{})_i`$ is the total (indistinguishable) number of $`\nu _\mu `$ and $`\overline{\nu }_\mu `$ induced muon events in the i-th energy bin and obs and th label the observed (i.e. simulated) and theoretical predicted muon event rates. With charge separation we calculate $`\chi _{\mu ^\pm }^2`$ separately for neutrinos and antineutrinos and use the sum $$\chi ^2=\chi _{\mu ^{},\nu }^2+\chi _{\mu ^{},\overline{\nu }}^2+\chi _{\mu ^+,\nu }^2+\chi _{\mu ^+,\overline{\nu }}^2.$$ (19) Next $`\chi ^2`$ is minimized as usual with respect to the parameters shown in the plots. We subtract the minimum value of $`\chi ^2`$ and define confidence levels according to the values of $$\mathrm{\Delta }\chi ^2=\chi ^2\chi _{min}^2$$ (20) This $`\mathrm{\Delta }\chi ^2`$ obeys a $`\chi ^2`$ distribution for $`k`$ degrees of freedom (i.e. the number of parameters fitted). The value of $`\mathrm{\Delta }\chi _{CL}^2`$ which corresponds to a given confidence level CL is given by: $$CL=_0^{\mathrm{\Delta }\chi _{CL}^2}\frac{x^{\frac{k}{2}1}e^{\frac{x}{2}}}{2^{\frac{k}{2}}\mathrm{\Gamma }(\frac{k}{2})}dx$$ (21) For two degrees of freedom $`(1\sigma ,2\sigma ,3\sigma )`$ corresponds to $`CL=(68.3\%,95.5\%,99.7\%)`$ and to $`\mathrm{\Delta }\chi _{CL}^2=(2.3,6.2,11.8)`$, as usual. We use in our analysis for the interval between 5 GeV and 20 GeV a total of 20 bins. We checked that the statistical significance of the results does not change much by changing the number of bins, as long as one has enough bins to describe the spectral information sufficiently. Finer binning does thus not improve the significance due to the statistical limitation by fluctuations. In this sense it is best to use around 20 energy bins which should not be a problem with the usual energy resolution of detectors. ## 6 Results As outlined above, we simulate a measurement with certain input parameters including statistical fluctuations for a given sign of $`\mathrm{\Delta }m_{31}^2`$. Then we try to fit this dataset, i.e. we re-extract the input parameters both with the right and the wrong sign of $`\mathrm{\Delta }m_{31}^2`$. Instead of a global fit to all parameters one can first fit the magnitude of $`|\mathrm{\Delta }m_{31}^2|`$ and $`\mathrm{sin}^22\theta _{23}`$ with good precision (a few percent ) by analyzing the muon spectrum for the dominant $`\nu _\mu \nu _\mu `$ and $`\overline{\nu }_\mu \overline{\nu }_\mu `$ disappearance oscillation<sup>5</sup><sup>5</sup>5One could combine the differential event rates coming from the $`\mu ^+`$ and $`\mu ^{}`$ beams, weighted to correct for the cross section differences, such that the matter dependence is almost removed. This allows a determination of $`|\mathrm{\Delta }m_{31}^2|`$ and $`\mathrm{sin}^22\theta _{23}`$ which is essentially independent of the sign of $`\mathrm{\Delta }m_{31}^2`$ and $`\mathrm{sin}^22\theta _{13}`$.. The following plots assume that this analysis is done and we analyze then with $`\mathrm{sin}^22\theta _{23}`$ fixed the matter induced effects and extract $`\mathrm{sin}^22\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$. We will see that this works very well if the mixing angle $`\theta _{13}`$ is not too small. The value of $`\mathrm{\Delta }\chi ^2`$ for the wrong sign of $`\mathrm{\Delta }m_{31}^2`$ will tell us the confidence level at which we can reject the incorrect sign of $`\mathrm{\Delta }m_{31}^2`$. For the right sign we can calculate confidence level contour lines from which we can read off the sensitivity for the determination of the relevant oscillation parameters. The $`\mathrm{\Delta }\chi ^2`$ values printed in the following figs. 5 \- 7 next to the CL contour lines are for the wrong sign of $`\mathrm{\Delta }m_{31}^2`$ relative to the best fit with the correct sign of $`\mathrm{\Delta }m_{31}^2`$. This are in figure 5 the values of the global minima with the wrong sign. In the other two figures we used the information on $`|\mathrm{\Delta }m_{31}^2|`$ from the fit in the $`\mathrm{sin}^22\theta _{23}`$-$`|\mathrm{\Delta }m_{31}^2|`$ plane to reject $`\mathrm{\Delta }m_{31}^2`$ values which are inconsistent with the $`|\mathrm{\Delta }m_{31}^2|`$ value obtained before. We therefore restrict the fit for the wrong sign of $`\mathrm{\Delta }m_{31}^2`$ to the local minimum of $`\chi ^2`$ in the neighborhood of the best fit of $`|\mathrm{\Delta }m_{31}^2|`$. Fig. 5 shows the $`1\sigma `$ and $`2\sigma `$ contour lines of the right sign fit for different $`|\mathrm{\Delta }m_{31}^2|`$ and different $`\theta _{13}`$ in the case with no charge separation (i.e. muons of either charge coming from $`\nu _\mu `$ or $`\overline{\nu }_\mu `$). With decreasing $`\theta _{13}`$, the precision in the determination of $`\mathrm{sin}^22\theta _{13}`$ gets worse, but this method allows to determine $`\mathrm{sin}^22\theta _{13}`$ already down to values of order $`𝒪(10^2)`$ (see also fig. 8 and discussion). Since the fitted spectrum includes the dominant disappearance channel, quite a good precision can be obtained also for $`|\mathrm{\Delta }m_{31}^2|`$. Charge separation capabilities improve the sensitivity for $`\mathrm{sin}^22\theta _{13}`$. This is shown in fig. 6, which shows the results of a fit to the appearance channel only (wrong sign muon events). In this case, which does not include the disappearance events in the analysis, one looses of course precision in finding the right value of $`|\mathrm{\Delta }m_{31}^2|`$. The precision for $`|\mathrm{\Delta }m_{31}^2|`$ can of course be improved in this case by a separate analysis of the disappearance spectrum . Fig. 6 shows that charge identification improves the precision in the determination of $`\mathrm{sin}^22\theta _{13}`$ for small values of $`\mathrm{sin}^22\theta _{13}`$ and the sensitivity will go down to values of order $`𝒪(10^310^4)`$. Note, however, that for not too small mixing angles $`\mathrm{sin}^22\theta _{13}`$ of about one magnitude below the current CHOOZ constraint, we find that the sum-rates without charge identification provide at least an equally well suited possibility to extract the value of $`\mathrm{sin}^22\theta _{13}`$. The determination of the sign of $`\mathrm{\Delta }m_{31}^2`$ is always better in the appearance channel, but note that the sum-rates give also quite good confidence levels for not too small mixing angles $`\theta _{13}`$. If charge identification is available then there is however an even better strategy by combining all available information, i.e. a global fit to the matter effects of both the appearance and disappearance channels. The results of this fit are shown in fig. 7. Now high precision is obtained both for $`\theta _{13}`$ and in the determination of $`\mathrm{\Delta }m_{31}^2`$. The presented fits were all done with a fixed mixing angle $`\mathrm{sin}^22\theta _{23}=1`$ and it is in principle no problem to perform a full three parameter fit for $`\mathrm{\Delta }m_{31}^2`$, $`\mathrm{sin}^22\theta _{13}`$ and $`\mathrm{sin}^22\theta _{23}`$. The contour lines of our fits have to be extended into the third $`\mathrm{sin}^22\theta _{23}`$ dimension and the precision is roughly given by the results in ref. . As explained above in the analytic discussion, there is no need to include further parameters like $`\mathrm{\Delta }m_{21}^2`$ and the CP-phase $`\delta `$ as long as the LMA MSW solution is not realized with a value of $`\mathrm{\Delta }m_{21}^2`$ close to its current upper limit of order $`𝒪(10^4)`$ . If this scenario were however realized, such that CP-effects play a role, then this would be another reason to go for measurements based on matter effects to largest possible baselines like $`7332`$ km. The point is that the relative magnitude of effects coming from $`\mathrm{\Delta }m_{21}^20`$ and $`\delta `$ decrease for longer baselines . Such a scenario would however also enable a measurement of the CP violating effects at shorter baselines such that two experiments, one at a large baseline (for matter effects) and another a shorter baseline (say 2800 km for CP violating effects) would be ideal. Fig. 8 shows finally the region of the $`\mathrm{\Delta }m_{31}^2`$-$`\mathrm{sin}^22\theta _{13}`$ parameter space where a determination of the sign of $`\mathrm{\Delta }m_{31}^2`$ will be possible. Shown are the $`1\sigma `$, $`2\sigma `$ and $`3\sigma `$ contour lines of a fit with the wrong sign of $`\mathrm{\Delta }m_{31}^2`$ for positive $`\mathrm{\Delta }m_{31}^2`$ (left plot) and negative $`\mathrm{\Delta }m_{31}^2`$ (right plot). For positive $`\mathrm{\Delta }m_{31}^2`$ the cross section favored neutrino channels show resonant matter effects whereas for negative $`\mathrm{\Delta }m_{31}^2`$ the cross section disfavored antineutrino channels are resonant. Thus in the case of positive $`\mathrm{\Delta }m_{31}^2`$ the confidence level at which the wrong sign can be excluded is slightly better. The dotted lines show the limits which can be obtained with charge identification from the spectral data of the appearance channels. The solid lines are based on an analysis of the combined channels without charge identification. Charge separation capabilities which allow to use the statistically favored pure appearance channel are thus very important if $`\mathrm{sin}^22\theta _{13}`$ turns out to be smaller than $`10^2`$. From this discussion it seems clear that one should include charge identification. But there are other issues which may turn out to be equally important as the discussion of statistics and rates which was presented here. If, for example, right sign muon rejection (i.e. charge identification) turns out to be less good than hoped, or if there were background issues which make it hard or impossible to isolate correctly the wrong sign $`\nu _e\nu _\mu `$ muon signal then our proposed method without charge identification should still work. Many of these issues are still under discussion . To illustrate the requirements for charge identification assume 100,000 muon events in a detector (which corresponds roughly to a baseline of $`3000`$ km and a beam energy of $`50`$ GeV with the usual beam and detector parameters) and right sign charge rejection of $`10^4`$, which would lead in average to 10 background events in the appearance channel which limits the ultimate sensitivity to $`\mathrm{sin}^22\theta _{13}`$ considerably. The problem can not be overcome by increasing statistics (i.e. by increasing $`N_{\mu ^\pm }N_{\text{kT}}ϵ_\mu `$) since the signal to background ratio stays constant. Any real detector may thus be limited by its right sign charge rejection capability and the number of right sign muons. This is also connected to the question which baseline should be used. The amount of background is typically directly proportional to the disappearance rates. This would favour larger baselines, since the disappearance rates drop faster than the appearance rates, improving the signal to background ratio for longer baselines. The decreased statistics in the $`\mathrm{sin}^22\theta _{23}`$ and $`|\mathrm{\Delta }m_{31}^2|`$ determination (with a precision of the order of a few percent) is not a problem since this will be limited by systematics and not statistics, even for the longest baselines. ## 7 Conclusions We studied in this paper different possibilities to use matter effects to determine the value and the sign of $`\mathrm{\Delta }m_{31}^2`$ and the magnitude of $`\mathrm{sin}^22\theta _{13}`$ in very long baseline experiments with neutrino factories. The analysis rests on the detection of muons coming from muon-neutrinos or antineutrinos and we distinguish detectors with and without right sign muon charge rejection capabilities. This may be important since the $`\nu _e\nu _\mu `$ and $`\overline{\nu }_e\overline{\nu }_\mu `$ appearance channels have rather small total event rates and they require very good efficiencies for the detection of wrong sign muon events. Backgrounds due to charge misidentification play an especially important role for very small $`\mathrm{sin}^22\theta _{13}`$ and at too short baselines, where the dominant non-oscillated neutrino rates are high. Studies of charge identification capabilities are presently performed . We discussed the relevant appearance and disappearance rates analytically and we performed numerical simulations which were used to test the parameter extraction from event rates in a statistical reliable way. Our results show for large baselines like $`7332`$ km that there are matter effects of comparable size in the event rates in the $`\nu _e\nu _\mu `$ and $`\overline{\nu }_e\overline{\nu }_\mu `$ appearance and in the $`\nu _\mu \nu _\mu `$ and $`\overline{\nu }_\mu \overline{\nu }_\mu `$ disappearance channels. Matter effects in the disappearance channels alone are however statistically somewhat disfavored, since they have to be extracted from the relatively large amount of un-oscillated events. Without sufficient charge identification capabilities there is however still the advantage that matter effects in the combined muon rates resulting from $`\nu _\mu `$ and $`\overline{\nu }_\mu `$, i.e. generated from the combined appearance and disappearance channels, add in a constructive way. This allows a determination of $`\mathrm{sin}^22\theta _{13}`$ and the sign of $`\mathrm{\Delta }m_{31}^2`$ even without charge identification down to mixings $`\mathrm{sin}^22\theta _{13}=10^2`$ and with a precision comparable to the appearance channels with perfect charge separation. This new method allows thus to measure or limit the mixing $`\mathrm{sin}^22\theta _{13}`$ roughly one order of magnitude below the present CHOOZ limit. With charge identification one can get better results from the appearance channels alone. We pointed however out that with charge identification the best results can be obtained by combining all information from the appearance and disappearance channels. In this case one can measure or limit the mixing down to $`\mathrm{sin}^22\theta _{13}10^4`$. We gave exclusion plots which display the parameter range where a determination of the sign of the mass squared difference $`\mathrm{\Delta }m_{31}^2`$ will be possible. Our results were obtained in the approximation where $`\mathrm{\Delta }m_{21}^2`$ corrections are negligible which allowed an analytic description. In the limit $`\mathrm{\Delta }m_{21}^2=0`$ CP violating effects drop out and matter effects can be understood in an effective two neutrino scheme which is inserted into the full three neutrino oscillation formulae. It should however be stressed that there will be sizable corrections to this picture for shorter baselines if $`\mathrm{\Delta }m_{21}^2`$ is at its upper limit. In this case the fit of $`\mathrm{\Delta }m_{31}^2`$, $`\mathrm{sin}^22\theta _{13}`$ and $`\mathrm{sin}^22\theta _{23}`$ had to include in addition $`\mathrm{\Delta }m_{21}^2`$ and the CP-phase $`\delta `$. These extra $`\mathrm{\Delta }m_{21}^20`$ effects become however much smaller for longer baselines, such that larger baselines are safer without loosing detection capabilities for matter effects. We did not discuss in more detail baseline and beam energy optimization since it depends on the preferred quantities, improvements on the knowledge of some parameters and the number of different baselines (i.e. beams and experiments) available. We mentioned however that a too short baseline could potentially also be dangerous if one neglects imperfect charge separation and other backgrounds. It seems therefore that larger baselines above 3000 km are also preferred from this point of view. We restricted our analysis therefore to a baseline of 7332 km which is very well suited for our study. We used a muon beam energy of $`20\mathrm{GeV}`$, since it gives a spectrum centered around the resonance energy of matter effects in earth and which seems to be preferred by recent studies of entry level neutrino factories . Acknowledgments: We wish to thank S.T. Petcov and A. Romanino for discussions on subjects related to this work.
warning/0004/hep-ph0004047.html
ar5iv
text
# SARGE: an algorithm for generating QCD-antennas ## Abstract We present an algorithm to generate any number of random massless momenta in phase space, with a distribution that contains the kinematical pole structure that is typically found in multi-parton QCD-processes. As an application, we calculate the cross-section of some $`e^+e^{}\text{partons}`$ processes, and compare SARGE’s performance with that of the uniform-phase space generator RAMBO. Considering that many multi-jet processes will occur in future hadron colliders, such as the LHC, it is necessary to calculate their cross-sections. A part of the amplitude of these processes consists of a multi-parton QCD-amplitude, and it is well known that the leading kinematic singularity structure of the squared matrix elements is given by the so-called antenna pole structure (APS). In particular, for $`n`$ gluons it is given by all permutations in the momenta of $$\frac{1}{(p_1p_2)(p_2p_3)(p_3p_4)\mathrm{}(p_{n1}p_n)(p_np_1)},$$ (1) where $`(p_ip_j)`$ denotes the Lorentz invariant scalar product of the gluon momenta $`p_i`$ and $`p_j`$. Actually, it is this kinematical structure that is implemented in algorithms based on the so called SPHEL approximation to calculate the amplitudes . But it is expected, and observed, that the same structure occurs in the exact matrix elements . For the integration of the differential cross-sections of the processes under consideration, the Monte Carlo method is the only option, and a phase space generator is needed. RAMBO is a robust and efficient algorithm to generate any number of random massless momenta in their center-of-mass frame (CMF) with a given energy. However, RAMBO generates the momenta distributed uniformly in phase space, so that a large number of events is needed to integrate integrands with the APS to acceptable precision. Especially when the evaluation of the integrand is time-consuming, which is the case for the exact matrix elements, this is highly inconvenient. In this paper, we introduce SARGE, an algorithm to generate any number of random massless momenta in their CMF with a given energy, distributed with a density that contains the APS. We shall show that it takes account for a substantial reduction in computing time in the calculation of cross-sections of multi-parton processes. We briefly sketch the outline of the SARGE-algorithm; a fuller discussion, appropriate to hadronic initial states as well, will be given elsewhere . The name SARGE stands for Staggered Antenna Radiation GEnerator, and is inspired by the structure of the algorithm. It consists of the repeated use of the basic antenna density for the generation of a momentum $`k`$, given two momenta $`p_1`$ and $`p_2`$: $$dA(p_1,p_2;k)=d^4k\delta (k^2)\theta (k^0)\frac{1}{\pi }\frac{(p_1p_2)}{(p_1k)(kp_2)}g\left(\frac{(p_1k)}{(p_1p_2)}\right)g\left(\frac{(kp_2)}{(p_1p_2)}\right).$$ (2) Here, $`g`$ is a function that serves to regularize the infrared and collinear singularities, as well as to ensure normalization over the whole space for $`k`$: therefore, $`g(\xi )`$ has to vanish sufficiently fast for both $`\xi 0`$ and $`\xi \mathrm{}`$. At this point, we take the simplest possible function we can think of, that has a sufficiently regularizing behavior. We introduce a positive non-zero number $`\xi _\text{m}`$ and take $$g(\xi )=\frac{1}{2\mathrm{log}\xi _\text{m}}\theta (\xi \xi _\text{m}^1)\theta (\xi _\text{m}\xi ),$$ (3) which forces the value of $`\xi `$ to be between $`\xi _\text{m}^1`$ and $`\xi _\text{m}`$, and is normalized such that $`𝑑A=1`$. Let us immediately adopt the notation $$\xi _1=\frac{(p_1k)}{(p_1p_2)}\text{and}\xi _2=\frac{(kp_2)}{(p_1p_2)}.$$ (4) The main motivation to make the regularizing function depend on $`\xi _1`$ and $`\xi _2`$ is that it makes $`dA`$ completely invariant under Lorentz-and scale transformations of the momenta. Consequently, the number $`\xi _\text{m}`$ gives a cut-off for the quotients $`\xi _1`$ and $`\xi _2`$ of the scalar products of the momenta, and not for the scalar products themselves. It is, however, possible to relate $`\xi _\text{m}`$ to the total energy $`\sqrt{s}`$ in the CMF and a cut-off $`s_0`$ on the invariant masses, i.e., the requirement that $$(p_i+p_j)^2s_0$$ (5) for all pairs of momenta $`p_ip_j`$. This can be done by choosing $$\xi _\text{m}=\frac{s}{s_0}\frac{(n+1)(n2)}{2},$$ (6) where $`n`$ is the total number of momenta. With this choice, the invariant masses $`(p_1+k)^2`$ and $`(k+p_2)^2`$ are regularized, but can still be smaller than $`s_0`$ so that the whole of the demanded phase space is covered. The $`s_0`$ can be derived from physical cuts $`p_T`$ on the transverse momenta and $`\theta _0`$ on the angles between the outgoing momenta: $$s_0=\mathrm{\hspace{0.33em}2}p_T^2\mathrm{min}(1\mathrm{cos}\theta _0,\left(1+\sqrt{1p_T^2/s}\right)^1).$$ (7) We now give the algorithm to generate $`k`$ under the basic antenna density. Let $`k^0`$, $`\varphi `$ and $`\theta `$ denote the absolute value, the polar angel and the azimuthal angle of $`\stackrel{}{k}`$ in the frame for which $`\stackrel{}{p}_1=\stackrel{}{p}_2`$ with $`\stackrel{}{p}_1`$ along the positive $`z`$-axis. To generate $`k`$, one should ###### Algorithm 1 (BASIC ANTENNA) 1. determine the direction of $`\stackrel{}{p}_1`$ in the CMF of $`p_1`$ and $`p_2`$; 2. generate two numbers $`\xi _1`$, $`\xi _2`$ independently, each from the density $`g(\xi )/\xi `$; 3. compute from these the values $`k^0`$ and $`\mathrm{cos}\theta `$; 4. generate $`\varphi `$ uniformly in $`[0,2\pi )`$; 5. construct the momentum $`k`$ in the CMF of $`p_1`$ and $`p_2`$; 6. boost the result to the actual frame in which $`p_1`$ and $`p_2`$ were given. The RAMBO algorithm was developed with the aim to generate the flat phase space distribution of $`n`$ massless momenta as uniformly as possible. The differential density is given by $$dV_n(\{p\})=\delta (\sqrt{s}P^0)\delta ^3(\stackrel{}{P})\underset{i=1}{\overset{n}{}}d^4p_i\delta (p_i^2)\theta (p_i^0),$$ (8) where $`P=_{i=1}^np_i`$. Let us denote $$dA_{j,k}^i=dA(q_j,q_k;q_i),\text{and}\xi _{k,l}^{i,j}=\frac{(p_ip_j)}{(p_kp_l)}.$$ (9) To include the APS in the density, one should ###### Algorithm 2 (QCD ANTENNA) 1. generate massless momenta $`q_1`$ and $`q_n`$ in CMF; 2. generate $`n2`$ momenta $`q_j`$ by the basic antennas $`dA_{1,n}^2dA_{2,n}^3dA_{3,n}^4\mathrm{}dA_{n2,n}^{n1}`$; 3. compute $`Q=_{j=1}^nq_j`$, and the boost and scaling transforms that bring $`Q^0`$ to $`\sqrt{s}`$ and $`\stackrel{}{Q}`$ to $`(0,0,0)`$; 4. for $`j=1,\mathrm{},n`$, boost and scale the $`q_j`$ accordingly, into the $`p_j`$. This way, the momenta $`p_j`$ are generated with differential density $`dV_n(\{p\})A_n^{\text{QCD}}(\{p\})`$, where $`A_n^{\text{QCD}}(\{p\})={\displaystyle \frac{s^2}{2\pi ^{n1}}}{\displaystyle \frac{g(\xi _{1,n}^{1,2})g(\xi _{1,n}^{2,n})g(\xi _{2,n}^{2,3})g(\xi _{2,n}^{3,n})\mathrm{}g(\xi _{n2,n}^{n2,n1})g(\xi _{n2,n}^{n1,n})}{(p_1p_2)(p_2p_3)(p_3p_4)\mathrm{}(p_{n1}p_n)(p_np_1)}}.`$ (10) We point out that, whereas the product $`dA_{1,n}^2\mathrm{}dA_{n2,n}^{n1}`$ contains a factor $`(p_1p_n)`$ in the numerator, the scaling transformation carries a Jacobian that is precisely $`s^2/(p_1p_n)^2`$, thus leading to a perfectly symmetric APS. Usually, the event generator is used to generate cut phase space. If a generated event does not satisfy the physical cuts, it is rejected. In the calculation of the weight coming with an event, the only contribution coming from the functions $`g`$ is, therefore, their normalization. In total, this gives a factor $`1/(2\mathrm{log}\xi _\text{m})^{2n4}`$ in the density. Because we are dealing with gluon momenta, we want to symmetrize the density. This can be done by re-labeling the momenta using a random permutation: ###### Algorithm 3 (SYMMETRIZATION) 1. generate a random permutation $`\sigma S_n`$ and put $`p_ip_{\sigma (i)}`$ for all $`i=1,\mathrm{},n`$. An algorithm to generate the random permutations can be found in . As a result, the differential density becomes $`dV_n(\{p\})\left({\displaystyle \frac{1}{n!}}{\displaystyle \underset{\text{perm.}}{}}A_n^{\text{QCD}}(\{p\})\right),`$ (11) where the sum is over all permutations of $`(1,\mathrm{},n)`$. An efficient algorithm to calculate a sum over permutations can be found in . When doing calculations with this algorithm on a phase space cut such that $`(p_i+p_j)^2>s_0`$ for all $`ij`$ and some reasonable $`s_0>0`$, we notice that a very high percentage of the generated events does not pass the cuts. An important reason why this happens is that the cuts, generated by the choices of $`g`$ (Eq. ​(3)) and $`\xi _\text{m}`$ (Eq. ​(6)), are implemented only on the variables $`\xi _{k,l}^{i,j}`$ that appear explicitly in the generation of the QCD-antenna. Therefore, an improvement is obtained as follows. Let $`𝐏_m`$ denote the subspace of $`[1,1]^m`$ for which $`|x_ix_j|1`$ for all $`i,j=1,\mathrm{},m`$, and let us denote the number of $`\xi _{k,l}^{i,j}`$-variables that has to be generated $`n_\xi =2n4`$. An improvement is obtained if the generation of these variables is replaced by ###### Algorithm 4 (IMPROVEMENT) 1. generate $`(x_1,\mathrm{},x_{n_\xi })`$ distributed uniformly in $`𝐏_{n_\xi }`$; 2. define $`x_0=0`$ and put, for all $`i=2,\mathrm{},n1`$, $$\xi _{i1,n}^{i1,i}e^{(x_{2i3}x_{2i4})\mathrm{log}\xi _\text{m}},\xi _{i1,n}^{i,n}e^{(x_{2i2}x_{2i4})\mathrm{log}\xi _\text{m}}.$$ (12) Because all the variables $`x_i`$ are distributed uniformly such that $`|x_ix_j|1`$, all quotients $`\xi _{k,l}^{i,j}`$ with $`(i,j)`$ and $`(k,l)`$ in $`\{(i1,i),(i,n)|i=2,\mathrm{},n1\}`$ are distributed such that they satisfy $`\xi _\text{m}^1\xi _{k,l}^{i,j}\xi _\text{m}`$. This is an improvement on the previous situation, because then only the quotients $`\xi _{i1,n}^{i1,1}`$ and $`\xi _{i1,n}^{i,n}`$ with $`i=2,\mathrm{},n1`$ satisfied the relation. In terms of the variables $`x_i`$, this means that the volume of $`𝐏_{n_\xi }`$ is generated, which is $`n_\xi +1`$, instead of the volume of $`[1,1]^{n_\xi }`$, which is $`2^{n_\xi }`$. We have to note here that this improvement only makes sense because there is a very efficient algorithm to generate the uniform distribution in $`𝐏_m`$ . The total density changes such that the product of the $`g`$-functions in Eq. ​(10) has to be replaced by $$g_{n2}^𝐏(\xi _\text{m};\{\xi \})=\frac{1}{(n_\xi +1)(\mathrm{log}\xi _\text{m})^{n_\xi }}\times \{\begin{array}{cc}1\hfill & \text{if }(x_1,\mathrm{},x_{n_\xi })𝐏_{n_\xi },\hfill \\ 0\hfill & \text{if }(x_1,\mathrm{},x_{n_\xi })𝐏_{n_\xi },\hfill \end{array}$$ (13) where the variables $`x_i`$ are functions of the variables $`\xi _{k,l}^{i,j}`$ as defined by $`(\text{12})`$. Again, only the normalization has to be calculated for the weight of an event. We compare SARGE with RAMBO in the calculation of the cross-section of the processes $$e^+e^{}\gamma ^{}q\overline{q}g,q\overline{q}q\overline{q},q\overline{q}q^{}\overline{q}^{},q\overline{q}gg,q\overline{q}q\overline{q}g,q\overline{q}q^{}\overline{q}^{}g,q\overline{q}ggg.$$ (14) The squared matrix element was calculated with the algorithm presented in , suitably adapted for these processes. We used massless electrons and quarks, and took the sum over final-state helicities and the average over initial-state helicities. We also summed over the color configurations of the final states. The center-of-mass energy $`\sqrt{s}`$ was fixed to $`500`$ GeV for the processes with $`5`$ outgoing momenta, and to $`100`$ GeV for the other processes. The cuts on the phase space where fixed with choices of a parameter $`\tau `$, which is related to the cut-off $`s_0`$ on the squares of the outgoing momenta (Eq. ​(5)) by $$s_0=\frac{2s\tau }{n(n1)},$$ (15) where $`n`$ is the number of outgoing momenta. If $`\tau =1`$, then $`s_0`$ is larger than the maximal value that is kinematically allowed. The couplings and charges in various processes were all set to the value $`1`$, since they only contribute a factor to the cross-section, which is irrelevant for this analysis. The results of the computer runs are given in the tables below. Presented are the final result for the cross-section $`\sigma `$ in units of $`\text{GeV}^2`$, the number of generated events $`N_{\text{ge}}`$, the number of accepted events $`N_{\text{ac}}`$, and the cpu-time consumed $`t_{\text{cpu}}`$ in seconds. All Monte Carlo runs were performed on a single 440-MHz UltraSPARC-IIi processor, and were stopped when an expected error of $`3\%`$ was reached. The final results for the cross-sections are irrelevant in our discussion, and are just printed to show that the results with SARGE and RAMBO are compatible within the $`3\%`$ error estimate. The most important conclusion that can be drawn from the results is that SARGE needs less accepted events than RAMBO for the given error estimate, especially for small values of $`\tau `$, i.e., for phase space that comes close to the singularities of the QCD-amplitudes. (Remember that the ratio of the volumes of cut phase space and whole phase space is given by $`N_{\text{ac}}/N_{\text{ge}}`$ for RAMBO.) As a result, less evaluations of the matrix elements have to be done which accounts for a large gain in computer time. It is true that SARGE is “ineffective” in the sense that many of the generated events have to be rejected because they do not satisfy the cuts imposed, but this is fully compensated by the fact that generating random numbers is much cheaper than evaluating matrix elements nowadays. For the last four processes, no results with RAMBO and $`\tau =0.01`$ are presented, but we observe that $`t_{\text{cpu}}>130,000`$ seconds. The fraction of phase space covered with five massless momenta and $`\tau =0.01`$ is $`0.893\pm 0.001`$. | $`e^+e^{}q\overline{q}g`$ | | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | | $`\sigma `$ | $`1.85\text{e -}5`$ | $`1.85\text{e -}5`$ | $`1.53\text{e -}4`$ | $`1.58\text{e -}4`$ | $`2.61\text{e -}4`$ | $`2.66\text{e -}4`$ | $`6.26\text{e -}4`$ | $`6.41\text{e -}4`$ | | $`N_{\text{ge}}`$ | $`7,691`$ | $`25,782`$ | $`10,777`$ | $`24,801`$ | $`10,806`$ | $`37,121`$ | $`11,437`$ | $`366,614`$ | | $`N_{\text{ac}}`$ | $`5,503`$ | $`6,536`$ | $`9,436`$ | $`20,112`$ | $`9,852`$ | $`33,577`$ | $`10,860`$ | $`359,447`$ | | $`t_{\text{cpu}}`$ | $`251`$ | $`293`$ | $`429`$ | $`899`$ | $`451`$ | $`1,503`$ | $`497`$ | $`16,124`$ | | $`e^+e^{}q\overline{q}q\overline{q}`$ | | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | | $`\sigma `$ | $`9.79\text{e -}9`$ | $`10.4\text{e -}9`$ | $`7.72\text{e -}7`$ | $`7.86\text{e -}7`$ | $`1.90\text{e -}6`$ | $`1.83\text{e -}6`$ | $`7.39\text{e -}6`$ | $`7.00\text{e -}6`$ | | $`N_{\text{ge}}`$ | $`64,384`$ | $`158,678`$ | $`32,492`$ | $`27,091`$ | $`34,701`$ | $`29,642`$ | $`41,744`$ | $`113,368`$ | | $`N_{\text{ac}}`$ | $`4,428`$ | $`4,551`$ | $`9,894`$ | $`15,328`$ | $`13,081`$ | $`22,297`$ | $`20,150`$ | $`107,021`$ | | $`t_{\text{cpu}}`$ | $`775`$ | $`786`$ | $`1,718`$ | $`2,606`$ | $`2,256`$ | $`3,778`$ | $`3,578`$ | $`18,038`$ | | $`e^+e^{}q\overline{q}q^{}\overline{q}^{}`$ | | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | | $`\sigma `$ | $`5.38\text{e -}9`$ | $`5.30\text{e -}9`$ | $`4.07\text{e -}7`$ | $`4.24\text{e -}7`$ | $`1.00\text{e -}6`$ | $`1.02\text{e -}6`$ | $`3.95\text{e -}6`$ | $`3.89\text{e -}6`$ | | $`N_{\text{ge}}`$ | $`98,840`$ | $`245,138`$ | $`50,052`$ | $`45,963`$ | $`63,398`$ | $`50,873`$ | $`71,254`$ | $`366,166`$ | | $`N_{\text{ac}}`$ | $`6,696`$ | $`7,022`$ | $`15,392`$ | $`25,883`$ | $`23,989`$ | $`38,145`$ | $`34,584`$ | $`345,323`$ | | $`t_{\text{cpu}}`$ | $`1,165`$ | $`1,198`$ | $`2,664`$ | $`4,346`$ | $`4,133`$ | $`6,434`$ | $`5,843`$ | $`58,708`$ | | $`e^+e^{}q\overline{q}gg`$ | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | | $`\sigma `$ | $`1.76\text{e -}7`$ | $`1.70\text{e -}7`$ | $`1.86\text{e -}5`$ | $`1.95\text{e -}5`$ | $`5.19\text{e -}5`$ | $`5.27\text{e -}5`$ | $`5.40\text{e -}4`$ | | $`N_{\text{ge}}`$ | $`96,942`$ | $`268,407`$ | $`42,321`$ | $`86,608`$ | $`50,552`$ | $`298,073`$ | $`50,414`$ | | $`N_{\text{ac}}`$ | $`6,579`$ | $`7,677`$ | $`12,945`$ | $`48,902`$ | $`19,091`$ | $`223,530`$ | $`26,551`$ | | $`t_{\text{cpu}}`$ | $`1,363`$ | $`1,597`$ | $`3,619`$ | $`6,398`$ | $`3,802`$ | $`43,913`$ | $`5,287`$ | | $`e^+e^{}q\overline{q}q\overline{q}g`$ | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | | $`\sigma `$ | $`2.04\text{e -}11`$ | $`1.91\text{e -}11`$ | $`4.05\text{e -}8`$ | $`4.08\text{e -}8`$ | $`1.68\text{e -}7`$ | $`1.61\text{e -}7`$ | $`1.48\text{e -}6`$ | | $`N_{\text{ge}}`$ | $`4,028,648`$ | $`4,017,888`$ | $`238,220`$ | $`97,035`$ | $`203,237`$ | $`210,325`$ | $`176,710`$ | | $`N_{\text{ac}}`$ | $`5,616`$ | $`5,094`$ | $`14,216`$ | $`33,239`$ | $`19,522`$ | $`121,734`$ | $`29,492`$ | | $`t_{\text{cpu}}`$ | $`4,530`$ | $`3,941`$ | $`10,333`$ | $`23,875`$ | $`14,159`$ | $`87,756`$ | $`21,407`$ | | $`e^+e^{}q\overline{q}q^{}\overline{q}^{}g`$ | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | | $`\sigma `$ | $`1.05\text{e -}11`$ | $`1.05\text{e -}11`$ | $`2.19\text{e -}8`$ | $`2.23\text{e -}8`$ | $`9.07\text{e -}8`$ | $`8.86\text{e -}8`$ | $`7.85\text{e -}7`$ | | $`N_{\text{ge}}`$ | $`5,596,725`$ | $`6,929,475`$ | $`436,225`$ | $`188,693`$ | $`377,384`$ | $`522,602`$ | $`305,426`$ | | $`N_{\text{ac}}`$ | $`7,730`$ | $`8,844`$ | $`26,154`$ | $`64,558`$ | $`36,042`$ | $`302,724`$ | $`51,044`$ | | $`t_{\text{cpu}}`$ | $`5,882`$ | $`6,494`$ | $`17,595`$ | $`43,104`$ | $`24,764`$ | $`201,801`$ | $`34,700`$ | | $`e^+e^{}q\overline{q}ggg`$ | | | | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | | $`\tau `$ | $`0.5`$ | | $`0.1`$ | | $`0.05`$ | | $`0.01`$ | | alg. | SARGE | RAMBO | SARGE | RAMBO | SARGE | RAMBO | SARGE | | $`\sigma `$ | $`1.31\text{e -}11`$ | $`1.30\text{e -}11`$ | $`3.63\text{e -}7`$ | $`3.54\text{e -}7`$ | $`1.63\text{e -}6`$ | $`1.54\text{e -}6`$ | $`1.85\text{e -}5`$ | | $`N_{\text{ge}}`$ | $`5,926,016`$ | $`6,650,538`$ | $`366,538`$ | $`131,617`$ | $`303,003`$ | $`186,257`$ | $`335,307`$ | | $`N_{\text{ac}}`$ | $`8,194`$ | $`8,475`$ | $`21,918`$ | $`45,157`$ | $`29,018`$ | $`107,897`$ | $`56,008`$ | | $`t_{\text{cpu}}`$ | $`7,407`$ | $`7,398`$ | $`18,120`$ | $`36,958`$ | $`24,036`$ | $`88,318`$ | $`46,673`$ | As an extra illustration, we also present the convergence to zero of the expected error during the Monte Carlo-run for a few cases. In Fig. ​1, we plot the relative error as function of the number of generated events using a double-log scale. We first of all observe that the curves for SARGE are less spiky, which shows that SARGE takes care for a substantial part of the singular behavior of the integrand. Every time a RAMBO-event hits a singularity, a term much larger than the average so far is added to the Monte Carlo sum, resulting in an increase of the expected error. Furthermore, we observe that the SARGE-error converges quicker than the RAMBO-error, except in the case of $`e^+e^{}q\overline{q}ggg`$ with $`\tau =0.05`$. However, this is a plot of the error as function of the number of generated events, and we know that many SARGE-events have to be rejected. A more realistic view is given by a plot of the error as function of cpu-time (Fig. ​2), which clearly shows that SARGE outperforms RAMBO.
warning/0004/quant-ph0004070.html
ar5iv
text
# Quantum dynamics and statistics of two coupled down-conversion processes ## 1 Introduction Optical parametric processes yield a wide variety of optical phenomena. It is not surprising that many new phenomena will arise if a parametric process is coupled to another one or to a different optical process. For instance, the superposition of signal photons originating from two down-convertors with aligned idler beams leads to nontrivial quantum interference effects . Parametric process coupled via Kerr interaction to an auxiliary mode, exhibiting quantum Zeno effect is another nice example . Many such composite systems (usually called nonlinear couplers) has thoroughly been studied in the literature. All-optical switching in the assymetric nonlinear coupler operating by the second-harmonic generation has been investigated in and its non-classical behaviour has been discussed in . The quantum dynamics and statistics of the symmetric coupler containing two second-harmonic processes have been examined in . The coupler composed of one linear waveguide and one nonlinear waveguide operating by the down-coversion process has been investigated in from the point of view of all optical switching. The occurrence of quantum Zeno and anti-Zeno effects in a similar device has been reported in . Amplitude behaviour of two linearly coupled down-conversion processes has been studied in . Short-length analysis of this device has been given in . In this paper we deal with interesting phenomena arising as a consequence of linear interaction between beams propagating through the symmetric nonlinear coupler, which is composed of two nonlinear waveguides based on the down-conversion processes. In fact this arrangement can be looked at as a continuous version of famous Mandel’s experiment , involving real physical interaction between the two down-conversion processes. In Section 2 the equations of motion are derived and their analytical solutions are given. Sections 3 and 4 are devoted to the study of quantum dynamics and statistics of the coupler. Its non-classical properties are discussed in Section 5. ## 2 Equations of motion and their solution The coupler which is investigated in this article is composed of two nonlinear waveguides operating by the down-conversion processes in a directional arrangement (see Fig. 1). The linear energy exchange by means of evanescent waves between pump, signal and idler beams is considered. The nonlinear media are assumed to be lossy. If such a system is far from resonance, the effective description involving only the field variables is adequate . The effective momentum operator then reads $$\widehat{G}=\widehat{G}_1+\widehat{G}_2+\widehat{G}_{res.}+\widehat{G}_{res.syst.}+\widehat{G}_{int.},$$ (1) where $`\widehat{G}_i`$ $`=`$ $`\mathrm{}{\displaystyle \underset{j=P_i,S_i,I_i}{}}k_j\widehat{a}_j^{}\widehat{a}_j+\mathrm{}\left(\mathrm{\Gamma }_i\widehat{a}_{P_i}\widehat{a}_{S_i}^{}\widehat{a}_{I_i}^{}+\text{h.c.}\right)\text{ for\hspace{1em}i=1,2},`$ $`\widehat{G}_{res.}`$ $`=`$ $`\mathrm{}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle \underset{j=P_i,S_i,I_i}{}}{\displaystyle \underset{l}{}}k_{lj}\widehat{b}_{lj}^{}\widehat{b}_{lj},`$ $`\widehat{G}_{res.syst.}`$ $`=`$ $`\mathrm{}{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle \underset{j=P_i,S_i,I_i}{}}{\displaystyle \underset{l}{}}\left(\kappa _{lj}\widehat{a}_j\widehat{b}_{lj}^{}+\text{h.c.}\right),`$ $`\widehat{G}_{int.}`$ $`=`$ $`\mathrm{}(\kappa _P\widehat{a}_{P_1}\widehat{a}_{P_2}^{}+\kappa _S\widehat{a}_{S_1}\widehat{a}_{S_2}^{}+\kappa _I\widehat{a}_{I_1}\widehat{a}_{I_2}^{}+\text{h.c.}),`$ (2) where $`\widehat{a}_j(\widehat{a}_{j}^{}{}_{}{}^{})`$, $`j=P_1,P_2,S_1,S_2,I_1,I_2`$ are annihilation (creation) operators of pump, signal, and idler modes. Corresponding wavevectors along the z-axis of propagation are $`k_{P_1}`$, $`k_{P_2}`$, $`k_{S_1}`$, $`k_{S_2}`$, $`k_{I_1}`$ and $`k_{I_2}`$. Linear coupling constants between pump, signal and idler modes are denoted $`\kappa _P`$, $`\kappa _S`$ and $`\kappa _I`$. Nonlinear coupling constants are denoted as $`\mathrm{\Gamma }_1`$ and $`\mathrm{\Gamma }_2`$. Each mode $`j`$ is coupled via linear coupling constant $`\kappa _{lj}`$ to the $`l`$-th reservoir mode characterized by annihilation (creation) operators $`\widehat{b}_{lj}(\widehat{b}_{lj}^{}{}_{}{}^{})`$ and wavevector $`k_{lj}`$ along z-axis of propagation. The symbol $`\mathrm{}`$ denotes the reduced Planck constant and h.c. represents Hermitian conjugate terms. The model represented by the momentum operator (1) is symmetric both under the exchange $`12`$, $`\kappa _P\kappa _P^{}`$, $`\kappa _S\kappa _S^{}`$, $`\kappa _I\kappa _I^{}`$ and under the exchange $`SI`$. Since the dynamical behaviour of the system is completely determined by its momentum operator, the symmetries are conserved during evolution. This is convenient because it is not necessary to write down all calculated quantities, the rest being simply obtained by the above mentioned exchanges. Substituting (1) into the Heisenberg equations of motion $`(i\mathrm{}\frac{d}{dz}\widehat{a}=[\widehat{G},\widehat{a}])`$, introducing slowly varying operators $`\widehat{A}_j(z)=\widehat{a}_j(z)\text{exp}(ik_jz)`$ and applying the Wigner-Weisskopf approximation , we arrive at the following Heisenberg-Langevin equations of motion $`{\displaystyle \frac{d\widehat{A}_{P_1}}{dz}}`$ $`=`$ $`\gamma _{P_1}\widehat{A}_{P_1}+i\kappa _P^{}\widehat{A}_{P_2}\text{exp}(i\mathrm{\Delta }k_Pz)+i\mathrm{\Gamma }_1^{}\widehat{A}_{S_1}\widehat{A}_{I_1}\text{exp}(i\mathrm{\Delta }l_1z)+\widehat{L}_{P_1}(z),`$ $`{\displaystyle \frac{d\widehat{A}_{S_1}}{dz}}`$ $`=`$ $`\gamma _{S_1}\widehat{A}_{S_1}+i\kappa _S^{}\widehat{A}_{S_2}\text{exp}(i\mathrm{\Delta }k_Sz)+i\mathrm{\Gamma }_1\widehat{A}_{P_1}\widehat{A}_{I_1}^{}\text{exp}(i\mathrm{\Delta }l_1z)+\widehat{L}_{S_1}(z),`$ where $`\mathrm{\Delta }k_k=k_{k_1}k_{k_2}`$, $`k=P,S,I`$ are linear mismatches, $`\mathrm{\Delta }l_i=k_{P_i}k_{S_i}k_{I_i}`$, $`i=1,2`$ are nonlinear mismatches, $`\gamma _j`$, $`j=P_1,P_2,S_1,S_2,I_1,I_2`$ are damping contants and the Langevin forces $`\widehat{L}_j(z)`$ are assumed to be Markoffian $`\widehat{L}_j(z)`$ $`=`$ $`\widehat{L}_j^{}(z)=\widehat{L}_j(z)\widehat{L}_k(z^{})=0,`$ $`\widehat{L}_j^{}(z)\widehat{L}_k(z^{})`$ $`=`$ $`2\gamma _jn_{dj}\delta _{jk}\delta (zz^{}),`$ $`\widehat{L}_j(z)\widehat{L}_k^{}(z^{})`$ $`=`$ $`2\gamma _j(n_{dj}+1)\delta _{jk}\delta (zz^{}).`$ (4) Here angle brackets denote the averaging over the reservoirs, $`n_{dj}`$ is one-mode mean photon number of the $`j`$-th reservoir, $`\delta _{jk}`$ is the Kronecker symbol and $`\delta (z)`$ is the Dirac delta function. It is useful to introduce the auxiliary quantities $$K_{S_i}=\gamma _{S_i}i\mathrm{\Delta }K_{S_i},K_{I_i}=\gamma _{I_i}+i\mathrm{\Delta }K_{I_i},i=1,2,$$ (5) where $$\mathrm{\Delta }K_{S_{1,2}}=\frac{\mathrm{\Delta }k\pm \mathrm{\Delta }k_S}{2},\mathrm{\Delta }K_{I_{1,2}}=\frac{\mathrm{\Delta }k\pm \mathrm{\Delta }k_I}{2}$$ (6) and $$\mathrm{\Delta }k=\frac{1}{2}\underset{i=1}{\overset{2}{}}(k_{S_i}+k_{I_i}k_{P_i}).$$ (7) The mismatch (7) contains wavevectors of all modes and thus characterizes the overall phase mismatch. This important quantity will be called global mismatch in the following. If we assume the pump modes $`P_1`$, $`P_2`$ are stimulated by the classical strong coherent fields $$\widehat{A}_{P_1}(z)\xi _{P_1}\text{exp}(i\mathrm{\Delta }k_Pz/2),\widehat{A}_{P_2}(z)\xi _{P_2}\text{exp}(i\mathrm{\Delta }k_Pz/2),$$ (8) the system of equations of motion, represented by (2), splits into two independent sets. The first one corresponds to $`\{\widehat{A}_{S_1},\widehat{A}_{S_2},\widehat{A}_{I_1}^{},\widehat{A}_{I_2}^{}\}`$ operators and the second one corresponds to their adjoints. In what follows we will confine ourselves to the first set. The special choice of the phases of classical amplitudes (8) leads, after the substitutions $$\widehat{A}_{S_1}(z)=\widehat{C}_{S_1}(z)\text{exp}(i\mathrm{\Delta }K_{S_1}z),\widehat{A}_{I_1}^{}(z)=\widehat{C}_{I_1}^{}(z)\text{exp}(i\mathrm{\Delta }K_{I_1}z),$$ (9) to the system of linear differential equations with constant coefficients of the form $`{\displaystyle \frac{d\widehat{C}_{S_1}}{dz}}`$ $`=`$ $`K_{S_1}\widehat{C}_{S_1}+i\kappa _S^{}\widehat{C}_{S_2}+iG_1\widehat{C}_{I_1}^{}+\widehat{}_{S_1}(z),`$ $`{\displaystyle \frac{d\widehat{C}_{I_1}^{}}{dz}}`$ $`=`$ $`K_{I_1}\widehat{C}_{I_1}^{}i\kappa _I\widehat{C}_{I_2}^{}iG_1^{}\widehat{C}_{S_1}+\widehat{}_{I_1}^{}(z),`$ (10) where $$\widehat{}_{S_1}(z)=\widehat{L}_{S_1}(z)\text{exp}(i\mathrm{\Delta }K_{S_1}z),\widehat{}_{I_1}^{}(z)=\widehat{L}_{I_1}^{}(z)\text{exp}(i\mathrm{\Delta }K_{I_1}z)$$ are modified Langevin forces, and $`G_1=\mathrm{\Gamma }_1\xi _{P_1}`$, $`G_2=\mathrm{\Gamma }_2\xi _{P_2}`$ are rescaled nonlinear coupling constants. The system of Eqs. (2) can be solved using the Laplace transformation method and method of variation of constants. Returning to the operators $`\widehat{A}_j`$, the solution can be written in the following matrix form $$\widehat{𝐀}(z)=𝐌(z)[𝐗(z)\widehat{𝐀}(0)+\widehat{𝐑}(z)],$$ (11) where we have introduced the vector \[$`()^T`$ means the transposition\] $$\widehat{𝐀}(z)=(\widehat{A}_{S_1}(z),\widehat{A}_{S_2}(z),\widehat{A}_{I_1}^{}(z),\widehat{A}_{I_2}^{}(z))^T,$$ (12) the vector of reservoir contribution $$\widehat{𝐑}(z)=(\widehat{R}_{S_1}(z),\widehat{R}_{S_2}(z),\widehat{R}_{I_1}(z),\widehat{R}_{I_2}(z))^T,$$ (13) the diagonal matrix of mismatches $$𝐌(z)=\text{diag}(\text{exp}(i\mathrm{\Delta }K_{S_1}z),\text{exp}(i\mathrm{\Delta }K_{S_2}z),\text{exp}(i\mathrm{\Delta }K_{I_1}z),\text{exp}(i\mathrm{\Delta }K_{I_2}z))$$ (14) and the matrix of coefficients $$X_{ij}(z)=\underset{k=1}{\overset{4}{}}(A_k)_{ij}\text{exp}(\lambda _kz)\text{for}i,j=1,..,4,$$ (15) where $`𝐀_k=[{\displaystyle \underset{ik=1}{\overset{4}{}}}(\lambda _k\lambda _i)]^1(\lambda _k^3𝐚+\lambda _k^2𝐛+\lambda _k𝐜+𝐝),k=1,..,4.`$ (16) Four-dimensional matrices $`𝐛`$, $`𝐜`$, $`𝐝`$ can be found in Appendix A and $`𝐚`$ is unity matrix. The quantities $`\lambda _k`$, $`k=1,..,4`$ in (15) and (16) are single roots of the polynomial $$\mathrm{\Delta }=x^4+ax^3+bx^2+cx+d,$$ (17) with coefficients $`a`$ $`=`$ $`\gamma _{S_1}+\gamma _{S_2}+\gamma _{I_1}+\gamma _{I_2},`$ $`b`$ $`=`$ $`L_S+L_I+\overline{L}_1+\overline{L}_2+K_{S_1}K_{I_2}+K_{I_1}K_{S_2},`$ $`c`$ $`=`$ $`L_SK_{I_2}+L_IK_{S_1}+\overline{L}_1K_{S_2}+\overline{L}_2K_{I_1}`$ $`+|\kappa _S|^2K_{I_1}+|\kappa _I|^2K_{S_2}|G_1|^2K_{I_2}|G_2|^2K_{S_1},`$ $`d`$ $`=`$ $`K_{S_1}K_{I_1}K_{S_2}K_{I_2}+|\kappa _S|^2K_{I_1}K_{I_2}+|\kappa _I|^2K_{S_1}K_{S_2}|G_1|^2K_{S_2}K_{I_2}`$ (18) $`|G_2|^2K_{S_1}K_{I_1}+|\kappa _S\kappa _IG_1^{}G_2|^2,`$ where $$L_j=K_{j_1}K_{j_2}+|\kappa _j|^2,j=S,I,\overline{L}_i=K_{S_i}K_{I_i}|G_i|^2,i=1,2.$$ If the roots of polynomial (17) are multiple, we can obtain the solution using the same methods. ## 3 Quantum dynamics To investigate the dynamical behaviour of the coupler, we have to find roots of the characteristic polynomial (17). If the damping is neglected and perfect phase matching is assumed, the polynomial is quadratic in $`x^2`$ and its roots are easy to find. In this case we can obtain periodical solution, exponentially amplifying solution or a combination of these two depending on the parameters of the process . If either the losses are included and all phase mismatches are zero, or losses are neglected and phase mismatches are retained, we arrive at the fourth-order polynomial with real coefficients. The roots can be found using the Cardan formulae; unfortunately they are of complicated form and it is more convenient to solve for the roots numerically. In the most general case we need to solve the fourth-order equation with complex coefficients, a task, which can only be performed with the help of a computer. However, there are certain physically realizable regimes $`\gamma _{S_1}`$ $`=`$ $`\gamma _{S_2}=\gamma _S,\gamma _{I_1}=\gamma _{I_2}=\gamma _I,`$ $`\mathrm{\Delta }k_S`$ $`=`$ $`\mathrm{\Delta }k_I=0,G_1=G_2{\displaystyle \frac{\kappa _S^{}|\kappa _I|}{\kappa _I|\kappa _S|}},`$ (19) for which the general polynomial (17) factorizes into two second-order ones. Their roots are $`\lambda _{1,2}`$ $`=`$ $`{\displaystyle \frac{\left[\gamma _S+\gamma _I+i\left(\left|\kappa _I\right|\left|\kappa _S\right|\right)\right]\pm \sqrt{\left[\gamma _S\gamma _I+i\left(\left|\kappa _S\right|+\left|\kappa _I\right|+\mathrm{\Delta }k\right)\right]^2+4\left|G_1\right|^2}}{2}},`$ $`\lambda _{3,4}`$ $`=`$ $`{\displaystyle \frac{\left[\gamma _S+\gamma _I+i\left(\left|\kappa _S\right|\left|\kappa _I\right|\right)\right]\pm \sqrt{\left[\gamma _S\gamma _I+i\left(\left|\kappa _S\right|+\left|\kappa _I\right|\mathrm{\Delta }k\right)\right]^2+4\left|G_1\right|^2}}{2}}.`$ Neglecting the damping ($`\gamma _S=\gamma _I=0`$), we can examine the influence of the global mismatch $`\mathrm{\Delta }k`$ to the dynamics of the coupler: 1. If $`\mathrm{\Delta }k(2|G_1|,2|G_1|)`$, then 1. for $`|\kappa _S|+|\kappa _I|(0,|\mathrm{\Delta }k|+2|G_1|)`$ all roots have the form $`\lambda _j=a_j+ib_j`$ with nonzero real and imaginary parts $`a_j`$ and $`b_j`$ 2. for $`|\kappa _S|+|\kappa _I|(|\mathrm{\Delta }k|+2|G_1|,+\mathrm{})`$ all roots are purely imaginary. 2. If $`\mathrm{\Delta }k(\mathrm{},2|G_1|)(2|G_1|,+\mathrm{})`$, then 1. for $`|\kappa _S|+|\kappa _I|(|\mathrm{\Delta }k|2|G_1|,|\mathrm{\Delta }k|+2|G_1|)`$ all roots are the same as in 1(a) 2. for $`|\kappa _S|+|\kappa _I|(0,|\mathrm{\Delta }k|2|G_1|)(|\mathrm{\Delta }k|+2|G_1|,+\mathrm{})`$ all roots are the same as in 1(b). Assuming the symmetrical linear coupling, $`|\kappa _S|=|\kappa _I|`$, the imaginary parts $`b_j`$ of $`\lambda _j`$ in cases 1(a) and 2(a) vanish and all roots acquire real values. In what follows, if $`a_j0`$ ($`a_j=0`$), we will say that the coupler operates in the hyperbolic (elliptic) regime. A more instructive demonstration of regions of different dynamical behaviour is given in Fig. 2. Notice first that $`\mathrm{\Delta }k`$ and its counterpart $`|\kappa _S|+|\kappa _I|`$ affect the process in symmetrical ways. Now let us look closely at the interplay between the linear coupling and global mismatch. If $`|\mathrm{\Delta }k|<2|G_1|`$ and the coupling strength is gradually increased, one can see that the coupler crosses the border between the hyperbolic regime (hatched area) and elliptic regime when a certain value of the linear coupling strength $`|\kappa _S|+|\kappa _I|`$ is attained. More interestingly, for large phase mismatch $`|\mathrm{\Delta }k|2|G_1|`$ (even such that almost no energy is converted from the pump mode to the signal and idler modes), the system moves up along the vertical line $`\mathrm{\Delta }k=\text{const.}`$ in Fig. 2, and it enters the region of instability characterized by the domination of the down-conversion part of the evolution when $`|\kappa _S|+|\kappa _I|=\mathrm{\Delta }k2|G_1|`$. For even stronger linear coupling the coupler leaves the region of instability again (crossing the line $`|\kappa _S|+|\kappa _I|=\mathrm{\Delta }k+2|G_1|`$), the oscillatory character of the evolution is restored and pump photons gradually cease to decay. Interpreting (somewhat loosely) the linear coupling as a kind of continuous measuring process, we can look at the just described behaviour as being a manifestation of the well-known Zeno or anti-Zeno effects . Also here the strong influence of the “measuring apparatus” leads to the hindering of the decay of the originally unstable system. On the contrary under certain conditions (here nonzero phase mismatch $`\mathrm{\Delta }k`$), the decay of the unstable system can be enhanced by a frequent (here continuous) monitoring of the unstable system . Our Fig. 2 clearly shows the competition between these two opposite tendencies. ## 4 Quantum statistics The quantum-statistical properties of the coupler are best studied employing the normal characteristic function containing complete statistical information about the system. The model represented by momentum operator (1) together with the linearization procedure (8) lead to the Gaussian characteristic function corresponding to the generalized superposition of coherent fields and quantum noise $`C_𝒩(\{\beta _j\},z)`$ $`=`$ $`\text{exp}\{{\displaystyle \underset{j=1}{\overset{4}{}}}[B_j(z)|\beta _j|^2+{\displaystyle \frac{1}{2}}(C_j(z)\beta _{j}^{}{}_{}{}^{2}+\text{c.c.})+`$ (21) $`+{\displaystyle \underset{k=1,j<k}{\overset{4}{}}}(D_{jk}(z)\beta _j^{}\beta _k^{}+\overline{D}_{jk}(z)\beta _j\beta _k^{}+\text{c.c.})`$ $`+(\beta _j\xi _j^{}(z)\text{c.c.})]\},`$ where the following identification $`S_11`$, $`S_22`$, $`I_13`$, $`I_24`$ has been done. The complex amplitudes $`\xi _1(z)`$, $`\xi _2(z)`$, $`\xi _3(z)`$, $`\xi _4(z)`$ are mean values of operators $`\widehat{A}_{S_1}(z)`$, $`\widehat{A}_{S_2}(z)`$, $`\widehat{A}_{I_1}(z)`$, $`\widehat{A}_{I_2}(z)`$, c.c. means the complex conjugated terms and $`B_j(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_j^{}(z)\mathrm{\Delta }\widehat{A}_j(z),C_j(z)=(\mathrm{\Delta }\widehat{A}_j(z))^2,`$ $`D_{jk}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_j(z)\mathrm{\Delta }\widehat{A}_k(z),\overline{D}_{jk}(z)=\mathrm{\Delta }\widehat{A}_j^{}(z)\mathrm{\Delta }\widehat{A}_k(z)`$ (22) for $`j,k=S_1,S_2,I_1,I_2`$ are noise functions. The complicated explicit expressions of the noise functions are given in Appendix B. The quantities $`C_j=C_j(0)`$ and $`B_j=B_j(0)+1`$ corresponding to the input beams are expressed under the condition of independence of incident beams in the form $$B_j=\text{cosh}^2(r_j)+n_{chj},C_j=\frac{1}{2}\text{exp}(i\theta _j)\text{sinh}(2r_j),$$ (23) where $`r_j`$ and $`\theta _j`$, $`j=1,2,3,4`$ are squeeze parameters and phases of the incident beams and $`n_{chj}`$ represents the mean number of external noise photons in the $`j`$-th mode. Assuming the unsqueezed input fields ($`r_j=0`$) the explicit expressions of noise functions (see Appendix B) lead to the following identities $`C_{S_1}(z)`$ $`=`$ $`C_{S_2}(z)=C_{I_1}(z)=C_{I_2}(z)=0,`$ $`D_{S_1S_2}(z)`$ $`=`$ $`\overline{D}_{S_1I_1}(z)=\overline{D}_{S_1I_2}(z)=0.`$ (24) The quantum-statistical properties of single and compound modes can be quantified by means of many statistical quantities. From these we will use the principal squeeze variance $`\lambda (z)`$ \[principal squeezing occurs if $`\lambda <1(2)`$ for single (compound) mode\], quadrature variances $`\left[\mathrm{\Delta }\genfrac{}{}{0pt}{}{\widehat{q}}{\widehat{p}}(z)\right]^2`$ \[quadrature squeezing occurs if $`(\mathrm{\Delta }\widehat{q})^2<1(2)`$ or $`(\mathrm{\Delta }\widehat{p})^2<1(2)`$ for single (compound) mode\], normal reduced factorial moments of the integrated intensity $`\frac{W^k(z)}{W(z)^k}1`$ \[they are negative for non-classical states, negative second moment reflects the sub-Poissonian photon statistics\] and the photon number distribution $`p(n,z)`$ \[quantum oscillations in $`p(n,z)`$ indicate the presence of state having no classical analogy\]. Adopting the standard definitions of the above mentioned quantities for single mode and using (4), it is straightforward to show that single modes do not exhibit any interesting behaviour. In particular non-classical light cannot develop from coherent inputs in single modes. It can arise only as a result of quantum correlations of modes. In the case of the compound mode $`(i,j)`$ the principal squeeze variance $`\lambda _{ij}(z)`$, quadrature variances $`[\mathrm{\Delta }\widehat{q}_{ij}(z)]^2`$, $`[\mathrm{\Delta }\widehat{p}_{ij}(z)]^2`$ and variance of the integrated intensity $`[\mathrm{\Delta }W_{ij}(z)]^2`$ are defined as follows $$\lambda _{ij}(z)=2\{1+B_i(z)+B_j(z)2\text{Re}[\overline{D}_{ij}(z)]|C_i(z)+C_j(z)+2D_{ij}(z)|\},$$ (25) $$(\mathrm{\Delta }\genfrac{}{}{0pt}{}{\widehat{q}}{\widehat{p}})^2=2\{1+B_i(z)+B_j(z)2\text{Re}[\overline{D}_{ij}(z)]\pm \text{Re}[C_i(z)+C_j(z)+2D_{ij}(z)]\}$$ (26) and $$[\mathrm{\Delta }W_{ij}(z)]^2=[\mathrm{\Delta }W_i(z)]^2+[\mathrm{\Delta }W_j(z)]^2+2\mathrm{\Delta }W_i(z)\mathrm{\Delta }W_j(z),$$ (27) where $`[\mathrm{\Delta }[W_i(z)]^2=B_i^2(z)+|C_i(z)|^2+2B_i(z)|\xi _i(z)|^2+2\text{Re}[C_i(z)\xi _{i}^{}{}_{}{}^{2}(z)]`$ (28) and $`\mathrm{\Delta }W_i(z)\mathrm{\Delta }W_j(z)`$ $`=`$ $`2\text{Re}[D_{ij}(z)\xi _i^{}(z)\xi _j^{}(z)\overline{D}_{ij}(z)\xi _i(z)\xi _j^{}(z)]`$ (29) $`+|D_{ij}(z)|^2+|\overline{D}_{ij}(z)|^2.`$ If the correlation function (29) is negative, we say that the corresponding modes are anti-correlated. The definitions of sum photon number distribution and $`k`$-th moment $`W_{ij}^k(z)`$ are rather complex and can be found in . ## 5 Discussion of results As we have already mentioned above, analytical expressions of required quantities are only available under certain simplifying and restrictive assumptions. Even in those cases the expressions are of a complicated form and thus almost useless for qualitative discussions. Therefore we will employ numerical methods. The analytical solutions, when available, may serve for checking the results of the numerical calculations. This section is devoted to the investigation of interesting phenomena arising from the linear coupling between two down-conversion processes. Each phenomenon is discussed in separate subsection. ### 5.1 Quadrature switching It was reported in that the symmetric coupler $`(|G_1|=|G_2|)`$ where only signal modes are linearly coupled $`(\kappa _I=0)`$ behaves as follows. If mode $`S_1`$ is squeezed in the given quadrature at the input, squeezing in a conjugated quadrature develops in mode $`S_2`$. Taking into account also linear exchange between idler modes, we can observe a similar phenomenon in quadratures of compound mode $`(S_1,I_1)`$. Let us assume that both down-conversion processes are spontaneous, linear coupling constants are symmetric $`\kappa _S=\kappa _I`$ and sufficiently strong. Changing now the phase $`\phi _{P_2}\text{arg}\xi _{P_2}`$ of the pump mode $`P_2`$ and leaving the phase $`\phi _{P_1}\text{arg}\xi _{P_1}`$ of the pump mode $`P_1`$ fixed, we can switch between quadratures at the output of mode $`(S_1,I_1)`$. Moreover, if the interaction length $`L`$ is appropriatelly chosen, squeezing in the given quadrature can be transferred to the conjugated one in a continuous way (see Fig. 3). ### 5.2 Linear coupling can compensate wrong phases It is well known that for small interaction lengths $`z`$ sub-Poissonian light can be generated in a nondegenerate down-conversion process in compound mode $`(S,I)`$ provided that the process is stimulated $`(\xi _S,\xi _I0)`$ and phases of incident beams fulfil the optimum phase condition $`\text{arg}(\xi _S\xi _I\xi _P^{})=\frac{\pi }{2}`$. On the other hand, if either the process is spontaneous or the phase condition is strongly violated, this mode is super-Poissonian. Let us assume that the process in the first waveguide is stimulated by amplitudes $`\xi _{S_1},\xi _{I_1}`$ strongly violating the optimum phase condition (say $`\text{arg}(\xi _{S_1}\xi _{I_1}\xi _{P_1}^{})=\frac{\pi }{2}`$) and the process in the second waveguide is spontaneous $`(\xi _{S_2}=\xi _{I_2}=0)`$. Introducing the linear coupling between the waveguides, the modes $`(S_1,I_1)`$ and $`(S_2,I_2)`$ can exhibit an interesting non-classical behaviour. The linear coupling restores the optimum phase condition and sub-Poissonian light is generated in mode $`(S_1,I_1)`$, surprisingly, for larger $`z`$ (see Fig. 4 (a)). Further, sub-Poissonian light is also generated in mode $`(S_2,I_2)`$ for small $`z`$ (Figures 4 (b) and 5), a phenomenon, which cannot be explained as easy as in the previous case. To deepen insight into this phenomenon we will resort to the analytical results. If losses are neglected, all mismatches are zero, $`\kappa \kappa _S=\kappa _I`$ is real, $`GG_1=G_2`$, and $`\kappa >|G|`$, we can calculate, using the results of Section 2 and (29), the correlation of fluctuations $`\mathrm{\Delta }W_{S_2}(z)\mathrm{\Delta }W_{I_2}(z)`$ $`=`$ $`|G|^2u^2(z)v^2(z)+2\kappa ^2|G||\xi _{S_1}||\xi _{I_1}|u(z)v^3(z)`$ (30) $`\times \text{sin}(\phi _P\phi _{S_1}\phi _{I_1}),`$ where $`\phi _P`$, $`\phi _{S_1}`$, $`\phi _{I_1}`$ are the phases of input coherent amplitudes $`\xi _P\xi _{P_1}=\xi _{P_2}`$, $`\xi _{S_1}`$, $`\xi _{I_1}`$ and $`u(z)`$ $`=`$ $`\text{cos}\left[\sqrt{2\left(\kappa ^2|G|^2\right)}z\right]{\displaystyle \frac{z}{4}}\sqrt{2\left(\kappa ^2|G|^2\right)}\text{sin}\left[\sqrt{2\left(\kappa ^2|G|^2\right)}z\right],`$ $`v(z)`$ $`=`$ $`{\displaystyle \frac{3\text{sin}\left[\sqrt{2\left(\kappa ^2|G|^2\right)}z\right]}{4\left[\sqrt{2\left(\kappa ^2|G|^2\right)}\right]}}+{\displaystyle \frac{z}{4}}\text{cos}\left[\sqrt{2\left(\kappa ^2|G|^2\right)}z\right].`$ (31) Since both the expressions on the right hand side (R.H.S.) of Eqs. (5.2) are real, only the second term on the R.H.S. of Eq. (30) can be negative depending on the sign of the product $`u(z)v(z)`$, and on the argument of sine function. Restricting ourselves to small $`z`$, we can expand $`u(z)`$ and $`v(z)`$ up to the $`z^3`$ around the origin, and approximate Eq. (30) by the expression $$\mathrm{\Delta }W_{S_2}(z)\mathrm{\Delta }W_{I_2}(z)|G|^2z^2+2\kappa ^2|G||\xi _{S_1}||\xi _{I_1}|z^3\text{sin}(\phi _P\phi _{S_1}\phi _{I_1}).$$ (32) Note first, that anti-correlation can only arise for sufficiently strong $`\kappa `$ and for sufficiently large $`z`$. It attains its maximum value if $`\phi _P\phi _{S_1}\phi _{I_1}=\text{arg}(\xi _P\xi _{S_1}^{}\xi _{I_1}^{})=\frac{\pi }{2}`$ (see Fig 4 (b)). It is also evident from the second term of the R.H.S. of Eq. (32) that the linear interaction enables us to affect the anti-correlation in mode $`(S_2,I_2)`$ via amplitudes $`\xi _{S_1}`$, $`\xi _{I_1}`$. Repeating the arguments leading to the formula (30) for mode $`(S_1,I_1)`$, we obtain $`\mathrm{\Delta }W_{S_1}(z)\mathrm{\Delta }W_{I_1}(z)`$ $`=`$ $`|G|^2\left[2\left(|\xi _{S_1}|^2+|\xi _{I_1}|^2\right)+1\right]u^2(z)v^2(z)`$ (33) $`2|G||\xi _{S_1}||\xi _{I_1}|\text{sin}(\phi _P\phi _{S_1}\phi _{I_1})`$ $`\times u(z)v(z)\left[u^2(z)+|G|^2v^2(z)\right].`$ Employing once more the expansion of (5.2) around $`z=0`$, we can see, that the second term on the R.H.S. of Eq. (33) is proportional to $`z`$ and thus its sign is given only by the argument of sine function. Our choice of the initial phases then implies that the contribution of the second term is positive in this approximation. However, the product $`u(z)v(z)`$ alternates and its amplitude increases with increasing $`z`$, suppressing the quantum noise represented by the first term on the R.H.S. of Eq. (33) and attaining the sub-Poissonian photon statistics for larger $`z`$. ### 5.3 Cross mode Up to now we have separately discussed modes localized either in the first or in the second waveguide. This subsection will be devoted to the investigation of non-classical behaviour occuring in cross mode $`(S_1,I_2)`$. The investigation of the down-conversion with strong pumping led to the conclusion that signal mode $`S`$ and idler mode $`I`$ do not exhibit any non-classical behaviour irrespectively of the fact, if they are spontaneous or stimulated by the coherent light . Obviously, mode $`(S_1,I_2)`$ compounded of modes $`S_1`$ and $`I_2`$ originating from two independent down-conversion processes cannot provide a non-classical light either. However, introducing the linear interaction between the processes, we can observe squeezing of vacuum fluctuations and sub-Poissonian photon statistics simultaneously in this mode (see Figure 6). To discover the origin of these phenomena we will employ once more the analytical solution. In the spirit of the derivation of Eq. (30) we can calculate the cross-correlation function $$D_{S_1I_2}(z)=\kappa |G|v^2(z),$$ (34) indicating, that linear exchange introduces the correlation between modes $`S_1`$ and $`I_2`$. This correlation reduces both vacuum fluctuations (see (25)) and fluctuations of photon number. To make the latter more clear, we can derive the following cross-correlation function up to $`z^3`$ $$\mathrm{\Delta }W_{S_1}(z)\mathrm{\Delta }W_{I_2}(z)2\kappa ^2|G||\xi _{S_1}||\xi _{I_1}|z^3\text{sin}(\phi _P\phi _{S_1}\phi _{I_1}).$$ (35) It is worth noting that the effect of noise reduction can be enhanced by increasing the amplitudes $`\xi _{S_1}`$ and $`\xi _{I_1}`$. ### 5.4 Mismatch-controlled switching Before discussing the last phenomenon we would like to mention several general remarks concerning the all-optical switching. This will enlight the motivation of the following discussion. Recent theoretical investigation of couplers has led to an interesting conclusion. Not only can they serve as a passive optical switchers, but they also provide the active control of the output of a particular waveguide by means of the input of the other one. There are at least two ways how to actively control the output beams. First, the coupling lenght of the coupler can be adjusted by changing the intensity of the strong classical input field . Second, the phase-controlled distribution of the quantum noise in couplers can be realized (see also Subection 5.1). There is, however, one more possibility how to affect the properties of the outgoing beams. Inspection of Fig. 2 reveals that one can change the dynamical behaviour of the beams by means of the global mismatch $`\mathrm{\Delta }k`$. Moreover, due to its global character (it contains all wavevectors), we can control one mode by means of another one, even though they directly do not interact. The following arrangement can illustrate this. Let us assume that both processes are spontaneous, nonzero wavevectors $`k_{S_1}`$, $`k_{S_2}`$, $`k_{I_1}`$, $`k_{I_2}`$ and $`k_{P_1}`$ are chosen in such a way, that they satisfy the matching conditions $`(\mathrm{\Delta }k_S=\mathrm{\Delta }k_I=\mathrm{\Delta }l_1=0)`$ and linear interaction is in operation. Now the increase of the $`z`$-th component $`k_{P_2}`$ of the wavevector of mode $`P_2`$ entails the inhibition of the decay of the pump photons in the first waveguide (see Figure 7). This phenomenon is easy to explain based on Fig. 2. Initially, the parameters of the coupler are such that the down-convertion part of the evolution is dominant (we are inside the hatched area). The global mismatch $`\mathrm{\Delta }k`$ decreases with increasing $`k_{P_2}`$ and the linear part of evolution grows dominant. Interpreting once again the linear interaction as a sort of continuous measurement, this effect can be looked at as a complementary effect to the Zeno or anti-Zeno-like effects described in Sec. 3. Unlike in Sec. 3 where initial condition (the value of $`\mathrm{\Delta }k`$) was kept constant and the strength of “measurement” was changed, here the strength of the “measurement” is the fixed quantity and the initial condition is continuously varied. In this way the influence of the “measurement” results in the speeding up the down-conversion for small values $`k_{P_2}`$ and slowing down the down-conversion for larger $`k_{P_2}`$. This corresponds to a transition from the anti-Zeno to Zeno regime. It is also worth noting, that the integrated intensity of mode $`(S_2,I_2)`$ depends on $`k_{P_2}`$ in the same way as mode $`(S_1,I_1)`$ does (see Fig. 7). This can be understood as follows. At the beginning (when $`k_{P_2}=0`$) the process in the first waveguide is perfectly matched $`(\mathrm{\Delta }l_1=0)`$ and the process in the second waveguide is strongly mismatched $`(\mathrm{\Delta }l_20)`$. The linear interaction, however, symmetrizes the device in the way that it partially mismatches the first process and partially compensates the mismatch in the second waveguide at the same time. ## 6 Conclusion The quantum dynamics and statistics of the symmetric nonlinear coupler operating by down-conversion process have been investigated. In a framework of strong pumping approximation we have solved analytically the Heisenberg-Langevin equations. The manifestation of Zeno and anti-Zeno effects has been demonstrated based on the analytical solution. The non-classical behaviour of beams involved has been studied based on numerical calculations. The phase-controlled redistribution of quantum noise between the quadratures can be achieved in mode $`(S_1,I_1)`$. The possibility of generation of sub-Poissonian light in modes $`(S_1,I_1)`$ and $`(S_2,I_2)`$ caused by the linear interaction of two super-Poissonian lights has been shown. Light exhibiting simultaneous squeezing of vacuum fluctuations and sub-Poissonian photon statistics can be obtained in cross mode $`(S_1,I_2)`$. The inhibition of the decay process in the first waveguide owing to the nonlinear matching of the second process has been observed. All these phenomena were shown to be robust against the presence of weak damping. ## Appendix A Matrices $`𝐛`$, $`𝐜`$, $`𝐝`$ of Eq. (16) $$𝐛=\left[\begin{array}{cccc}K_{I_1}+K_{S_2}+K_{I_2}& i\kappa _S^{}& iG_1& 0\\ i\kappa _S& K_{S_1}+K_{I_1}+K_{I_2}& 0& iG_2\\ iG_1^{}& 0& K_{S_1}+K_{S_2}+K_{I_2}& i\kappa _I\\ 0& iG_2^{}& i\kappa _I^{}& K_{S_1}+K_{S_2}+K_{I_1}\end{array}\right],$$ $`𝐜=[\begin{array}{ccc}K_{I_1}K_{S_2}+L_I+\overline{L}_2& i\kappa _S^{}\left(K_{I_1}+K_{I_2}\right)& iG_1\left(K_{S_2}+K_{I_2}\right)\\ i\kappa _S\left(K_{I_1}+K_{I_2}\right)& K_{S_1}K_{I_2}+L_I+\overline{L}_1& \kappa _I^{}G_2\kappa _SG_1\\ iG_1^{}\left(K_{S_2}+K_{I_2}\right)& \kappa _S^{}G_1^{}\kappa _IG_2^{}& K_{S_1}K_{I_2}+L_S+\overline{L}_2\\ \kappa _SG_2^{}\kappa _I^{}G_1^{}& iG_2^{}\left(K_{S_1}+K_{I_1}\right)& i\kappa _I^{}\left(K_{S_1}+K_{S_2}\right)\end{array}`$ $`\begin{array}{c}\kappa _IG_1\kappa _S^{}G_2\\ iG_2\left(K_{S_1}+K_{I_1}\right)\\ i\kappa _I\left(K_{S_1}+K_{S_2}\right)\\ K_{I_1}K_{S_2}+L_S+\overline{L}_1\end{array}],`$ $`𝐝=[\begin{array}{ccc}\overline{L}_2K_{I_1}+\left|\kappa _I\right|^2K_{S_2}& i\kappa _S^{}L_Ii\kappa _IG_1G_2^{}& iG_1\overline{L}_2+i\kappa _S^{}\kappa _I^{}G_2\\ i\kappa _SL_Ii\kappa _I^{}G_1^{}G_2& \overline{L}_1K_{I_2}+\left|\kappa _I\right|^2K_{S_1}& \kappa _I^{}G_2K_{S_1}\kappa _SG_1K_{I_2}\\ iG_1^{}\overline{L}_2i\kappa _S\kappa _IG_2^{}& \kappa _S^{}G_1^{}K_{I_2}\kappa _IG_2^{}K_{S_1}& \overline{L}_2K_{S_1}+\left|\kappa _S\right|^2K_{I_2}\\ \kappa _SG_2^{}K_{I_1}\kappa _I^{}G_1^{}K_{S_2}& iG_2^{}\overline{L}_1i\kappa _S^{}\kappa _I^{}G_1^{}& i\kappa _I^{}L_S+i\kappa _SG_1G_2^{}\end{array}`$ $`\begin{array}{c}\kappa _S^{}G_2K_{I_1}+\kappa _IG_1K_{S_2}\\ iG_2\overline{L}_1+i\kappa _S\kappa _IG_1\\ i\kappa _IL_S+i\kappa _S^{}G_1^{}G_2\\ \overline{L}_1K_{S_2}+\left|\kappa _S\right|^2K_{I_1}\end{array}].`$ ## Appendix B Noise functions $`B_{S_1}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}^{}(z)\mathrm{\Delta }\widehat{A}_{S_1}(z)={\displaystyle \underset{j=1}{\overset{4}{}}}\left(|X_{1j}|^2B_j+2\gamma _jn_{dj}\chi _{1j}\right)`$ $`+{\displaystyle \underset{j=1}{\overset{2}{}}}\left(2\gamma _{j+2}\chi _{1j+2}|X_{1j}|^2\right),`$ $`C_{S_1}(z)`$ $`=`$ $`(\mathrm{\Delta }\widehat{A}_{S_1}(z))^2={\displaystyle \underset{j=1}{\overset{2}{}}}\left(X_{1j}^2C_j+X_{1j+2}^2C_{j+2}^{}\right)\text{exp}(2i\mathrm{\Delta }K_{S_1}z),`$ $`D_{S_1S_2}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}(z)\mathrm{\Delta }\widehat{A}_{S_2}(z)={\displaystyle \underset{j=1}{\overset{2}{}}}\left(X_{1j}X_{2j}C_j+X_{1j+2}X_{2j+2}C_{j+2}^{}\right)`$ $`\times \text{exp}(i\mathrm{\Delta }k),`$ $`D_{S_1I_1}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}(z)\mathrm{\Delta }\widehat{A}_{I_1}(z)=[{\displaystyle \underset{j=1}{\overset{4}{}}}(X_{1j}X_{3j}^{}B_j+2\gamma _jn_{dj}\chi _{jj}^{13})`$ $`+{\displaystyle \underset{j=1}{\overset{2}{}}}(2\gamma _j\chi _{jj}^{13}X_{1j+2}X_{3j+2}^{})]\text{exp}[i(\mathrm{\Delta }K_{S_1}+\mathrm{\Delta }K_{I_1})z],`$ $`D_{S_1I_2}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}(z)\mathrm{\Delta }\widehat{A}_{I_2}(z)=[{\displaystyle \underset{j=1}{\overset{4}{}}}(X_{1j}X_{4j}^{}B_j+2\gamma _jn_{dj}\chi _{jj}^{14})`$ $`+{\displaystyle \underset{j=1}{\overset{2}{}}}(2\gamma _j\chi _{jj}^{14}X_{1j+2}X_{4j+2}^{})]\text{exp}[i(\mathrm{\Delta }K_{S_1}+\mathrm{\Delta }K_{I_2})z],`$ $`\overline{D}_{S_1S_2}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}^{}(z)\mathrm{\Delta }\widehat{A}_{S_2}(z)=[{\displaystyle \underset{j=1}{\overset{4}{}}}(X_{1j}^{}X_{2j}B_j+2\gamma _jn_{dj}\chi _{jj}^{21})`$ $`+{\displaystyle \underset{j=1}{\overset{2}{}}}(2\gamma _{j+2}\chi _{j+2j+2}^{21}X_{1j}^{}X_{2j})]\text{exp}(i\mathrm{\Delta }k_Sz),`$ $`\overline{D}_{S_1I_1}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}^{}(z)\mathrm{\Delta }\widehat{A}_{I_1}(z)={\displaystyle \underset{j=1}{\overset{2}{}}}\left(X_{1j}^{}X_{3j}^{}C_j^{}+X_{1j+2}^{}X_{3j+2}^{}C_{j+2}\right)`$ $`\times \text{exp}[i(\mathrm{\Delta }K_{S_1}\mathrm{\Delta }K_{I_1})z],`$ $`\overline{D}_{S_1I_2}(z)`$ $`=`$ $`\mathrm{\Delta }\widehat{A}_{S_1}^{}(z)\mathrm{\Delta }\widehat{A}_{I_2}(z)={\displaystyle \underset{j=1}{\overset{2}{}}}\left(X_{1j}^{}X_{4j}^{}C_j^{}+X_{1j+2}^{}X_{4j+2}^{}C_{j+2}\right)`$ (40) $`\times \text{exp}[i(\mathrm{\Delta }K_{S_1}\mathrm{\Delta }K_{I_2})z],`$ where $`X_{ij}=X_{ij}(z)`$ are defined in Eq. (15) and $`\chi _{ij}`$ $`=`$ $`\chi _{ij}(z)={\displaystyle _0^z}|X_{ij}(zz^{})|^2𝑑z^{},`$ $`\chi _{jl}^{ik}`$ $`=`$ $`\chi _{jl}^{ik}(z)={\displaystyle _0^z}X_{ij}(zz^{})X_{kl}^{}(zz^{})𝑑z^{}.`$ The rest of the noise functions can be obtained using the symmetry of the model. Acknowledgments We would like to thank J. Peřina Jr. for help with numerical calculations. Support by Grant No. VS96028, Research project CEZ:J14 ”Wave and Particle Optics” of the Czech Ministry of Education and Grant No. 202/00/0142 of Czech Grant Agency is acknowledged.
warning/0004/astro-ph0004071.html
ar5iv
text
# Unbiased reconstruction of the mass function using microlensing survey data. ## 1 Introduction. The success of the microlensing experiments has been impressive, several collaborations have reported the detection of microlensing amplification of stars, OGLE (Udalski et al. 2000), MACHO (Alcock et al. 1998), EROS (Derue et al. 1999)), DUO (Alard & Guibert 1997), VATT (Uglesich et al. 1999). In particular, the recent release by the OGLE II collaboration of more than 200 microlensing events is of great interest. One of the obvious promises of such data sets is the possibility to explore the mass function near the low mass end. Although, the relation between the mass function and the microlensing observations is not straightforward. In the classical scheme of analysis, one must first estimate the duration of the events by fitting the theoretical light curves, then compute the lensing rates, and finally relate these lensing rates to the mass function. The trouble is that fitting the theoretical light curve to the microlensing data to obtain the duration of the event is a highly degenerated process (Alard 1997, Han 1997, Wozniak & Paczyǹski 1997). Due to the source confusion in crowded fields, it is almost impossible to estimate the flux of the amplified source reliably. The flux of the source is severely biased by blending with neighboring sources. In such case, an over-estimation of the baseline flux will result in an under-estimation of the duration of the event. This bias is severe for unresolved stars and will result in a reduction in the estimation of the duration of a factor of 2,3, or even more. Since the bias on the mass of the lens goes like the square of the bias on the duration, the relevant bias on the mass function will be very large. It was already shown (Alard 1997, Han 1997) that the contribution of these unresolved, highly blended stars to the lensing rated was dominant at short durations. The consequence is that the estimation of the mass function at low mass end will be largely over-estimated, suggesting the existence of a large number of brown dwarf that actually do not exists. It is obvious that as long as the bias due to the blended sources has not been solved, any correct estimation of the mass function is impossible. This paper proposes a solution to this problem. We will see that using differential photometry obtained using the image subtraction method (Alard & Lupton 1998, Alard 2000) it is possible to estimate the mass function, even without knowing the baseline flux of the source. This method assumes that the distribution and kinematics of the lenses are known with good accuracy. It is important to emphasize that the uncertainties related to the structure of the Galaxy in the bulge region are several order of magnitude smaller than the uncertainties due to the blending bias. Many good model of the kinematic and structure of the bulge of our Galaxy are already available (Zhao, 1996, Fux 1997, Bissantz et al., 1997, Binney, Gerhard, & Spergel 1997) ## 2 The Method. ### 2.1 Introduction When analyzing microlensing observations one has to deal with the light curves of a number N of microlensing candidates. We will assume that we have purely differential photometry only, with a baseline flux equal to zero. We emphasize that high quality differential photometry can always be obtained after the events have been detected by classical methods by using the image subtraction method (Alard 1999). These light curves will be represented by the symbol $`𝐒_𝐢(𝐭)`$, ($`j=1,..,N`$). Assuming that un-amplified flux of the source is A, the expression of the $`𝐒_𝐢(𝐭)`$ as a function of time will be given by the theoretical microlensing amplification formula: $$S_j(t)=(A1)\frac{u(t)^2+2}{u(t)\sqrt{u(t)^2+4}}$$ (1) With: $$u(t)=\sqrt{u_0^2+\left(\frac{t}{t_E}\right)^2}$$ (we recall that we are working in the assumption that we have at our disposal a differential flux only, with for convenience a baseline flux equal to 0). The most important issue is of course how do we relate these observations to the microlensing parameters: the impact parameter, $`𝐮_\mathrm{𝟎}`$, and the time to cross the Einstein ring: $`𝐭_𝐄`$. And furthermore how do we relate the microlensing parameters to the mass function itself. The first serious difficulty we encounter is that of course $`𝐮_\mathrm{𝟎},𝐭_𝐄`$, cannot be extracted easily by fitting the light curve due to the extreme blending degeneracy (Alard 1997, Han 1997, Wozniak & Paczynski 1997). Thus our first and most essential step will be to derive a better set of parameters. This set of parameters will have to be non degenerated with respect to blending. The most simple and most natural way to deal with such problem is to express the observations as a linear combination of a small number of vectors. The best way to make such a decomposition is known as principal components analysis, it is equivalent to an eigen value decomposition. Thus all we have to do is to look for the firsts eigen functions, and to calculate the scalar product of the components with our vectors (time series) of observations. ### 2.2 Normalization Before making the principal components analysis we have to consider that one of the parameters, the amplitude of the magnified sources does not contain any information about the lensing event. Actually the reason of the the degeneracy of the fit is precisely the unknown amplitude of the source. The only relevant and meaningful parameters are $`𝐮_\mathrm{𝟎}\mathrm{and}𝐭_𝐄`$. Thus it is important to derive an estimator which is independent of the amplitude. Since the amplitude is a linear parameter, it is sufficient to normalize the vector of observations in order to get rid of the amplitude. This normalization can be made in the sense of the scalar product we are going to use to make our principal components decomposition. In such case, the normalized light curve $`\overline{S}_i(t)`$ can be expressed as: $$\overline{S}_i(t)=\frac{1}{\alpha }\times \frac{u(t)^2+2}{u(t)\sqrt{u(t)^2+4}}$$ And: $$\alpha =\sqrt{\left[\frac{u(t)^2+2}{u(t)\sqrt{u(t)^2+4}}\right]^2𝑑t}$$ Note that $`\overline{A}(t)`$ does not depend any more on the amplitude, but only on $`𝐮_\mathrm{𝟎}\mathrm{and}𝐭_𝐄`$. The most simple way to apply the previous normalization to the observations $`S_j`$ is to calculate directly the modulus of $`|S_j|`$ from the observations. Although it is obvious that calculating directly $`|S_i|`$ from the observations may not be optimal. To estimate $`|S_i|`$ we will use the following approach: first we fit a microlensing model to the observations. Even if the fit is very degenerated, one can always find a good solution which fits the data well. We will call this fitted model $`\stackrel{~}{S}_i`$. Once we get this light curve solution we estimate using the following formula: $$|S_i|=\sqrt{\stackrel{~}{S_i}(t)^2𝑑t}$$ We will use the same approach for all our other calculations, in the same way we will estimate the principal components not by taking directly the data, but by taking the models we have fitted to the light curves. And finally the scalar products with the principal components will be also estimated by cross products with the theoretical models. ### 2.3 Principal components As we have already explain, we will first fit a microlensing model $`\stackrel{~}{S}_i`$ to the light curves, normalize this series of N vectors to obtain the series of vectors $`\overline{\stackrel{~}{S}_i}`$, and then calculate the principal components of these $`\overline{\stackrel{~}{S}_i}`$ vectors. Usually it is possible to express almost all the information with a reduced number of principal components with the following linear decomposition: $$\overline{\stackrel{~}{S}_i}=\underset{j}{}a_{ij}P_j(t)$$ Where $`a_{ij}`$ is the projection of time series number i, $`\stackrel{~}{𝐒_𝐢}(𝐮_\mathrm{𝟎},𝐭_𝐄,𝐭)`$ on the principal component number j: $$a_{ij}(\mathrm{u}_0,\mathrm{t}_\mathrm{E})=\overline{\stackrel{~}{\mathrm{S}_\mathrm{i}}}(\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{t})\mathrm{P}_\mathrm{j}(\mathrm{t})\mathrm{dt}$$ Note that $`a_{ij}`$ is a function of $`𝐮_\mathrm{𝟎}\mathrm{and}𝐭_𝐄`$: $`a_{ij}=\beta (\mathrm{u}_0,\mathrm{t}_\mathrm{E})`$ For instance a typical set of 100 microlensing light curves can be expressed as the combination of only 4 components with an accuracy good to 1 %. The other principal components contains very little additional information, but mostly noise. ### 2.4 Statistical distribution of the projections on the principal components. We consider that most of the information is contained in a number $`\mathrm{N}_\mathrm{C}`$ of principal components. Thus almost all the information can be extracted from the distribution of the projection of the $`N_1`$ time series on the $`\mathrm{N}_\mathrm{C}`$ principal components. The first meaningful quantity concerning the distribution of the $`a_{ij}`$ is the first order moment of the distribution: $$a_j=\underset{i}{}a_{ij}$$ The second interesting quantity is the second order moment of the distribution: $$a_j^2=\underset{i}{}a_{ij}^2$$ But it is important to notice, that $`a_{ij}^2=\left[\beta (\mathrm{u}_0,\mathrm{t}_\mathrm{E})\right]^2`$, can be written as a scalar product of the data with a given function. Since we can always find a function $`F(\mathrm{t})`$ such that: $$a_{ij}^2=\left[\beta (\mathrm{u}_0,\mathrm{t}_\mathrm{E})\right]^2=\overline{\stackrel{~}{S_i}}(\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{t})\mathrm{F}(\mathrm{t})\mathrm{dt}$$ (2) Proof: To prove that $`F(t)`$ exists we have to show that Eq. (2) can be satisfayed for any point in the space of the parameters $`(\mathrm{u}_0,\mathrm{t}_\mathrm{E})`$. We can map this parameter space by using a regular grid with a step as small as we like. This grid will have an almost infinite number of rows ($`N_u`$) and columns ($`N_t`$). The total number of points in the grid is: $`N_G=N_u\times N_t`$. We define $`F(t)`$ by sampling this function in $`N_G`$ points within the limits of the parameter space of the variable $`t`$. Since $`N_G`$ is very large, one can always approximate Eq. (2) with the following formulae: $$a_{ij}^2=\left[\underset{m}{}S_i(u_0^k,t_E^l,t_m)F_m\right]\times \mathrm{\Delta }T$$ The above equation can be written $`N_G`$ times (for all the points in the grid). Since we have also $`N_G`$ unknown values $`F_m`$, we can write a full system of $`N_G`$ linear equations, which will allow to find the $`N_G`$ values $`F_m`$ which represents the function $`F(t)`$. Thus the function F(t) exists for all values of $`\mathrm{u}_0\mathrm{and}\mathrm{t}_\mathrm{E}`$. Using the principal components decomposition of $`\overline{\stackrel{~}{S_i}}`$ we can write: $$a_{ij}^2\underset{j=1}{\overset{\mathrm{N}_\mathrm{C}}{}}\left[P_j(\mathrm{t})\mathrm{F}(\mathrm{t})\mathrm{dt}\right]a_{ij}$$ Thus, finally we see that the distribution of the second moment is not more than a linear combination of the the $`\mathrm{N}_\mathrm{C}`$ firsts moments (mean). Thus the mean of the second moment will not bring any additional information with respect to the mean. With a similar reasoning it is possible to show that the same property is true for the $`N^{th}`$ moment. Consequently we can conclude that all the information concerning the distribution of the $`a_{ij}`$ is contained in the mean of these components. No additional information (uncorrelated information) will be found by looking at higher order moments. ### 2.5 From the principal components to the mass function. It is possible to re-express the previous definition of $`a_j`$ by using the number density $`\rho (\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{M})`$ to observe a given amplification of parameters $`(\mathrm{u}_0,\mathrm{t}_\mathrm{E})`$ with a lens of mass M. In such case, the sum can be approximated very closely with an integral expression: $$a_j\rho (\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{M})\mathrm{a}_\mathrm{j}(\mathrm{u}_0,\mathrm{t}_\mathrm{E})\mathrm{du}_0\mathrm{dt}_\mathrm{E}\mathrm{dM}$$ If we define the efficiency function of the experiment, $`\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})`$, the lensing rates for 1 solar mass lenses, $`\mathrm{\Gamma }_0(t)`$ and the mass function, $`\varphi (M)`$ the formulae for $`\rho `$ reads: $$\rho (\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{M})=\mathrm{\Gamma }_0\left(\frac{t_E}{\sqrt{M}}\right)\varphi (M)\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})$$ Leading to the following expression for $`a_j`$: $$a_j\mathrm{\Gamma }_0\left(\frac{t_E}{\sqrt{M}}\right)\varphi (M)\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})a_j(\mathrm{u}_0,\mathrm{t}_\mathrm{E})\mathrm{du}_0\mathrm{dt}_\mathrm{E}\mathrm{dM}$$ It is easy to re-arrange this integral in the following way: $$a_j\psi (M)\varphi (M)𝑑M$$ With: $$\psi (M)=\mathrm{\Gamma }_0\left(\frac{t_E}{\sqrt{M}}\right)\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})a_{ij}(\mathrm{u}_0,\mathrm{t}_\mathrm{E})\mathrm{du}_0\mathrm{dt}_\mathrm{E}$$ (3) Thus we see that basically projecting the microlensing light curve on a basis of function and taking the statistical sum is equivalent to projecting the mass function itself on another function. Consequently, we see that a projection in the space of the observation (the light curve) is directly equivalent to a projection in the space of the mass function. The problem of finding the optimal set of projections for the observations, is solved by using the principal component method. Then all we have to do is to calculate the “image” of these principal components in the space of the mass function. ## 3 MONTE-CARLO SIMULATIONS. ### 3.1 Introduction. To illustrate this method and show how the mass function can be reconstructed, we will use a series of Monte-Carlo simulations. We will simulate microlensing events by selecting the parameters of the events according to the density distribution, $`\rho (\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{M})`$. One additional parameter we will have to select is the amplitude of the source. To get random variables which reproduces the distribution $`\rho `$ we need to decompose the problem. Basically $`\rho (\mathrm{u}_0,\mathrm{t}_\mathrm{E},\mathrm{M})`$ has 3 parts: \- The rates, $`\mathrm{\Gamma }_M(t_E)`$: $$\mathrm{\Gamma }_M(t_E)=\mathrm{\Gamma }_0(\frac{t_E}{\sqrt{M}})$$ For $`\mathrm{\Gamma }_0`$ we will adopt the analytical expression given in Mao & Paczyǹski (1996). The efficiency function, $`\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})`$: For the efficiency we will adopt the following criteria: an event is detected if it has a minimum number ($`N_m`$) of data points above a 5 $`\sigma `$ threshold. For a given duration $`t_E`$ this criteria is simply equivalent to put a threshold on the impact parameter $`u_0`$. All we have to do is to search for the maximum value $`u_T`$ of $`u_0`$ such that the amplification of at least $`N_m`$ data points is above 5 $`\sigma `$. If the sampling is even, with a time step $`\delta t`$, it just mean that the amplification must be larger than 5 $`\sigma `$ in a window of time around the maxima $`\mathrm{\Delta }T=N_m\delta t`$. Provided the maxima is at the origin of the time axis, it will finally result in the condition: $$S_j(\mathrm{u}_\mathrm{T},\mathrm{t}_\mathrm{E},\frac{\mathrm{\Delta }\mathrm{T}}{2})>5\sigma $$ (4) The distribution of $`u_0`$ itself is uniform, to get et set of $`u_0`$ values we will use a random generator which will provide a uniform distribution between 0 and the maximum value for $`u_0`$, $`u_M(t_E)`$. $`u_M(t_E)`$ corresponds to the impact parameter for the brightest source given by Eq xx. Once we have selected $`t_E`$ from the distribution $`\mathrm{\Gamma }_M(t_E)`$, we will calculate $`u_{0,thresh}`$ using Eq xx, if our random $`u_0`$ value is above $`u_{0,thresh}`$, $`\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})=0`$, otherwise $`\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})=1`$. To summarize: $$\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})=\{\begin{array}{cc}1\hfill & u_0\text{ }<\text{ }u_T\hfill \\ & \\ 0\hfill & u_0\text{ }>\text{ }u_T\hfill \end{array}$$ The mass function, $`\varphi (M)`$ : In all our simulation we will adopt a pure power law expression for the mass function, with a lower ($`M_I`$) and an upper cut-off ($`M_S`$). $$\varphi (M)=\{\begin{array}{cc}M^\alpha \hfill & M_I\text{ }<\text{ M }<\text{ }M_S\hfill \\ & \\ 0\hfill & \text{otherwise}\hfill \end{array}$$ The amplitude: we will assume that the the flux of the un-amplified source amplitude distribution is a power law with an exponent of -2 (Zhao, et al. 1995), in a given range of amplitude $`A_{max}`$ and $`A_{min}`$. In our simulation, the amplitudes will be generated by Monte-Carlo method, using this power law. ### 3.2 Description of the implementation. To implement our Monte-Carlo simulation we will select the random variables in the following order: \- The amplitude of the source, according to the probability law $`A^\nu `$ \- The mass of the lens, according to the probability law $`A^\alpha `$ \- The duration of the event $`t_E`$ using the distribution, $`\mathrm{\Gamma }_M(t_E)`$ \- The impact parameter $`u_0`$ using an uniform distribution in the range, 0, $`u_M(t_E)`$. \- Once $`t_e`$ and $`u_0`$ are selected we apply the efficiency cut-off, using the function $`\mathrm{\Theta }(\mathrm{u}_0,\mathrm{t}_\mathrm{E})`$. This procedure produces a complete set of variables, $`A`$, $`t_E`$, $`u_0`$ for the events passing the efficiency cut-off. Form the variable we compute the microlensing light curve using Eq. 1. We add noise to the light curves according to the Poisson statistics. Using this procedure, it is possible to simulate a set of microlensing light curves $`S_j`$. ### 3.3 Example In this example we will take an amplitude range which is typical for microlensing images towards the Galactic Bulge: $`A_{max}=10^6`$, $`A_{min}=10`$ For the mass function we take the following parameters: $$\varphi (M)=\{\begin{array}{cc}M^2\hfill & \text{0.05 }<\text{ M }<\text{ 10.0}\hfill \\ & \\ 0\hfill & \text{otherwise}\hfill \end{array}$$ Using these parameters we simulated several series of microlensing light curves. For illustration, we extracted a sample of light curves from one of these simulations, they are presented in Fig. 1. The parameters were adjusted in order that in a simulation the total number of microlensing events simulated be close to 100. Then we proceeded to the principal components decomposition. The orthonormal set of principal components was calculated using a singular value decomposition. The first 4 principal components are presented in Fig. 2. The functions $`\psi (M)`$ corresponding to the principal components in the space of the mass function can be computed using Eq. 3. Projecting the data on the principal components is equivalent to projecting the mass function on $`\psi (M)`$. To illustrate our discussion, an example of $`\psi (M)`$ functions calculated using the settings of the previous Monte-Carlo simulation are presented in Fig. 3. ### 3.4 Estimating the mass function. To illustrate the ability of the method to reconstruct the mass function we will assume that exponent of the mass the mass function, $`\alpha `$ is unknown. For each simulation we will calculate the 4 principal component of the set of microlensing light curves we have simulated. We will compute the sum of the projection on a principal component of all the light curves $`<a_j>`$. Using Eq. 3 we will calculate the function which corresponds to each principal components in the space of the mass function. Then all we have to do is to compare the projection of a trial mass function with $`<a_j>`$. The trial mass function which match the 4 $`<a_j>`$ as close as possible (in a least-square tens) will be the best mass function. This procedure was applied for 1000 simulation, each time the program derived the best value of the exponent of $`\alpha `$. The histogram of the values of alpha for the 1000 simulations is presented in Fig. 4.
warning/0004/nucl-th0004063.html
ar5iv
text
# FAU–TP3–00/6 nucl-th/0004063 Goldberger–Treiman Relation and 𝑔_{𝜋⁢𝑁⁢𝑁} from the Three Quark BS/Faddeev Approach in the NJL Model ## 1 Introduction The Nambu-Jona-Lasinio(NJL) model is a quantum field theoretical effective quark model based on QCD. While the confinement, which is one of the most important properties of the low energy QCD, is not incorporated, it is the simplest quantum field theoretical quark model respecting the chiral symmetry and providing us with explicit examples how the spontaneous breaking of the chiral symmetry plays an important role in hadronic phenomena, in particular, the interpretation of pion, K, and $`\eta `$ as Nambu-Goldstone modes. Based on these advantages, not only mesons but also baryons have been extensively studied in the NJL model. The studies of baryons are mainly classified into the following two categories: (C1) the relativistic mean field approach, and (C2) the relativistic three-quark Bethe-Salpeter(3qBS )/Faddeev approach. Since the NJL model is a second quantized field theory, it is, in principle, possible to consider the effects of $`q\overline{q}`$ excitations in the nucleon. This is one of the most interesting targets in the studies of the nucleon structure beyond the non-relativistic constituent quark models. In this respect, on the one hand, the meanfield approach incorporates $`q\overline{q}`$ effects through the mesonic hedgehog fields leading to the solitonic picture of the nucleon, where, however, the manifest Lorentz covariance is unfortunately missing. On the other hand, the 3qBS /Faddeev approach, which respects the manifest Lorentz covariance leading to the quark-diquark picture of baryons, approximates the 3qBS kernel according to the ladder truncation scheme, which prevents us from studying non-trivial $`q\overline{q}`$ effects beyond the excitations of the “three-quark RPA vacuum”(the backward diagrams included in the ladder approximation). It is thus necessary to extend the 3qBS kernel in order to study the non-trivial $`q\overline{q}`$ effects in a Lorentz covariant manner, and therefore attempts have been (and are) made to go beyond the ladder truncation scheme . To extend the NJL 3qBS kernels, chiral symmetry imposes important constraints. In particular, it would be very useful to have a criterion which tells us which kinds of kernels lead to the PCAC relation<sup>1</sup><sup>1</sup>1We mean by “PCAC relation” in the sense of Eq.(34). correctly. One of the main aims of this paper is to derive such a criterion and to provide explicit examples of its application to some of the existing 3qBS kernels. We will first review how matrix elements of quantum one-body operators are evaluated systematically in the 3qBS framework by introducing classical external fields as a technical tool. Note that the evaluation of the bound state matrix elements in the 3qBS framework is not obvious beyond the diagramatic argument. We will provide a direct formula to evaluate the matrix element in terms of the Faddeev framework. We then consider which properties of the 3qBS kernel are required to satisfy the PCAC relation correctly. We consider the relevant Feynman diagrams to calculate the matrix elements for several 3qBS kernels, and apply our criterion to these 3qBS kernels. Although we are going to restrict our attention only to the PCAC case, these considerations themselves can be extended to other cases straightforwardly such as the electromagnetic current, the isospin current, the baryon number current, etc —actually, the PCAC case is the most complicated one<sup>2</sup><sup>2</sup>2The proof of the electromagnetic Ward identity in the Faddeev approach to the NJL model in the ladder truncation scheme can be found in ref., which was applied to the quark-diquark model in . Although these authors did not use the external field method, the results are consistent with our formalism. A systematic approach using the external field method can be found in ref... Even if the truncation scheme is consistent with the PCAC relation, the UV regularization schemes, which have been adopted so far in practical numerical calculations in the NJL model, usually, spoil it. It is thus worth while to evaluate $`g_{\pi NN}`$ and $`g_A`$ in the chiral limit in order to estimate explicitly the violation of the Goldberger-Treiman relation due to the UV regularization scheme. We will see that with the Euclidean sharp cut-off the violation is up to 4 %. The extracted value of $`g_{\pi NN}`$ is 13.2. We will also estimate the PCAC violation due to the cutoff scheme off the chiral limit, which is again up to 4%. We will evaluate the off-shell $`g_{\pi NN}(\stackrel{~}{g}_{\pi NN})`$ with the single-pole dominance approximation for the form factor $`h_A(q^2)`$. We obtain a very reasonable value $`\stackrel{~}{g}_{\pi NN}=13.5`$. However $`g_A`$ is 6 % larger than the experimental one. We will also consider the effect of the non-vanishing current quark mass on the Goldberger-Treiman relation, and find that it is in the opposite direction than expected from the experimental values. We will analyze this by including also the effect of the $`q\overline{q}`$ interaction in the color singlet iso-vector axial-vector channel, providing an analytic expression for the deviation from the Goldberger-Treiman relation by assuming that the vacuum is approximated according to the mean-field approximation. We will find that the effect of this additional $`q\overline{q}`$ channel works into the unwanted direction. We will see that, to resolve this problem, it is necessary either to improve the gap equation for the vacuum beyond the mean field approximation or to evaluate the on-shell $`g_{\pi NN}`$. ## 2 Bound State Matrix Elements and the Criterion (Sufficient Condition) for the PCAC Relation We consider effective quark Lagrangians with global $`SU(2)_f\times SU(3)_c`$ symmetries: $$=\overline{\psi }\left(i\overline{)}m_0\right)\psi +_\mathrm{I},$$ (1) where $`m_0`$ is the current quark mass, and $`\psi `$ and $`\overline{\psi }`$ are the quark bispinor fields. $`_\mathrm{I}`$ is a local, chirally symmetric four-fermionic interaction Lagrangian of NJL type, i.e., $`_\mathrm{I}={\displaystyle \underset{\mathrm{\Gamma }}{}}g_\mathrm{\Gamma }\left(\overline{\psi }\mathrm{\Gamma }\psi \right)^2`$, where $`\mathrm{\Gamma }`$ is a matrix with Dirac, isospin, and color indices. Examples are the original NJL type $`_I=g\left((\overline{\psi }\psi )^2(\overline{\psi }\gamma _5\stackrel{}{\tau }\psi )^2\right)`$ and the color current interaction type $`_I=g\left(\overline{\psi }\gamma ^\mu \frac{\lambda ^a}{2}\psi \right)^2`$, etc. (For detail, see Appendix B.1.) Since interaction Lagrangians of this type can all be dealt with in the same manner, we will not confine ourselves to a particular one. To evaluate a matrix element in the 3qBS framework is not really straight forward. This is because what is directly obtained in the 3qBS framework is not the nucleon eigen-ket $`|N`$ but the 3qBS amplitude<sup>3</sup><sup>3</sup>3The 3qBS amplitude is often referred to as the “wave function” due to historical reasons. However, we prefer to call it as “3qBS amplitude” to avoid unnecessary confusions., which itself is a matrix element of the type $`0|T\psi \psi \psi |N`$. We might thus suspect that a huge amount of information, which is contained in the ket vector $`|N`$, could be missing in the 3qBS amplitude. In the relativistic Faddeev framework, the situation is slightly more complicated —there are no immediate representations of the Faddeev amplitude in terms of the canonical operator formalism<sup>4</sup><sup>4</sup>4The Faddeev amplitude provides the same amount of informations as the 3qBS amplitude does, as we show in Appendix A.. Therefore, we will first review how to evaluate a matrix element in the 3qBS framework based on the canonical operator formalism. To this end, it is convenient to introduce space-time dependent external fields: $$v_\mu (x)=v_\mu ^a(x)\frac{\tau ^a}{2},a_\mu (x)=a_\mu ^a(x)\frac{\tau ^a}{2},m(x)=s(x)+ip^a(x)\gamma _5\frac{\tau ^a}{2},$$ (2) where $`v_\mu ^a(x)`$ is an isovector vector field, $`a_\mu ^a(x)`$ is an isovector axial-vector field, $`s(x)`$ is a scalar isoscalar field, and $`p^a(x)`$ is a pseudo-scalar isovector field. (The iso-spin index $`a`$ runs over 1, 2, 3.) The following Lagrangian density characterizes how these external fields couple to the associated quantized one-body operators: $`^{[e]}`$ $`=`$ $`\overline{\psi }\left(i\overline{)}\overline{)}v\overline{)}a\gamma _5m\right)\psi +_\mathrm{I}`$ $`=`$ $`\overline{\psi }i\overline{)}\psi +_Iv_\mu ^aV^{a\mu }a_\mu ^aA^{a\mu }\overline{\psi }m\psi ,`$ where $`V^{a\mu }\overline{\psi }\gamma ^\mu {\displaystyle \frac{\tau ^a}{2}}\psi `$ and $`A^{a\mu }\overline{\psi }\gamma ^\mu \gamma _5{\displaystyle \frac{\tau ^a}{2}}\psi `$. The superscripts $`[e]`$ or $`[0]`$ will be used to indicate quantities in the presence or absence, respectively, of the external fields $`v,a,mm_0`$. For simplicity, we include the current quark mass $`m_0`$ in the definition of the external field $`m`$. We assume that these external fields, $`v,a,m`$, are localized in space-time, i.e., they vanish except for some finite space-time region. After performing the necessary manipulations, we will eventually set $`a(x)v(x)0`$ and $`mm_0`$ in the whole space-time. This limit will be refereed to as the “vanishing external fields limit”. The 3qBS equation in the presence of the external fields is derived from the Schwinger-Dyson(SD) equation by means of the same technique as for vanishing external fields: $$G^{[e]}=G_0^{[e]}+G_0^{[e]}V^{[e]}G^{[e]},$$ (4) where $`V^{[e]}`$ is the sum of 2PI and 3PI interactions, and $`G_0^{[e]}`$ is an anti-symmetric combination of products of three constituent quark propagators in the presence of the external fields. For simplicity, we denote $$K^{[e]}G_0^{[e]}V^{[e]}.$$ (5) To avoid cumbersome notations, we often adopt the operator notation to suppress the explicit integration symbol $`{\displaystyle d^4x_1d^4x_2d^4x_3}`$, the explicit space-times coordinates $`(x_1,x_2,x_3)`$ and the explicit indices of the quark fields. Since the external fields are localized in space–time, a nucleon state “propagates” freely in the “past”. Therefore, $`G^{[e]}`$ has an asymptotic initial nucleon pole. The residue at this nucleon pole satisfies the following homogeneous 3qBS equation in the presence of the external fields: $$\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}=K^{[e]}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]},$$ (6) where $`\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}`$ is the 3qBS amplitude for the nucleon in the presence of the external fields with the isospin $`a`$, helicity $`\alpha `$ and asymptotic four momentum $`P=(E_N(\stackrel{}{P}^2),\stackrel{}{P})`$ with $`E_N(\stackrel{}{P}^2)=\sqrt{m_N^2+\stackrel{}{P}^2}`$ ($`m_N`$ : nucleon mass). In terms of the canonical operator formalism, it is expressed as: $`\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}(x_1,x_2,x_3)`$ (7) $`=`$ $`\begin{array}{c}0|\text{}T\psi (x_1)\psi (x_2)\psi (x_3)\hfill \\ \times \mathrm{exp}i{\displaystyle d^4x\left[v_\mu ^a(x)V^{a\mu }(x)+a_\mu ^a(x)A^{a\mu }(x)+\overline{\psi }\left(m(x)m_0\right)\psi \right]\text{}|N_\alpha ^a(\stackrel{}{P})}.\hfill \end{array}`$ (10) $`|\text{}0`$ and $`|\text{}N_\alpha ^a(\stackrel{}{P})`$ are the vacuum and the nucleon state vectors, respectively. They are eigenstates of the “unperturbed” Hamiltonian $`H^{[0]}`$ obtained by setting $`v=a=0,m=m_0`$. We adopt the covariant normalizations: $$0|\text{}0=1,N_\alpha ^{}^a^{}(\stackrel{}{P}^{})|\text{}N_\alpha ^a(\stackrel{}{P})=\frac{E_N(\stackrel{}{P}^2)}{m_N}(2\pi )^3\delta ^{(3)}(\stackrel{}{P}^{}\stackrel{}{P})\delta _{\alpha ^{}\alpha }\delta _{a^{}a}.$$ (11) Note that quantized operators are represented in the “interaction picture” in Eq.(7), where the “unperturbed Hamiltonian”<sup>5</sup><sup>5</sup>5 Note that $`H^{[0]}`$ is the full Heisenberg Hamiltonian in the absence of the external fields. is $`H^{[0]}`$ and the “perturbing Hamiltonian” is $$H_Id^3x\left(v_\mu ^a(x)V_{a\mu }(x)+a_\mu ^a(x)A^{a\mu }(x)+\overline{\psi }(x)\left(m(x)m_0\right)\psi (x)\right).$$ (12) We next consider the matrix element $`N|A_\mu ^b|N`$. Applying the functional derivative $`{\displaystyle \frac{i\delta }{\delta a_\mu ^a(x)}}`$ to both sides of Eq.(6), we obtain the following relation in the vanishing external field limit, $$\left(\frac{i\delta \mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}=\left(\frac{i\delta K^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[0]}+K^{[0]}\left(\frac{i\delta \mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}.$$ (13) The expressions for the “3qBS amplitudes” entering in Eq.(13) are given in terms of the canonical operator formalism as follows: $`\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[0]}(x_1,x_2,x_3)`$ $`=`$ $`0\left|\text{}T\psi (x_1)\psi (x_2)\psi (x_3)\right|N_\alpha ^a(\stackrel{}{P})`$ (14) $`\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{P})]^{[0]}(x_1,x_2,x_3)`$ $`=`$ $`N_\alpha ^a(\stackrel{}{P})\left|\text{}T\overline{\psi }(x_1)\overline{\psi }(x_2)\overline{\psi }(x_3)\right|0`$ $`\left({\displaystyle \frac{i\delta \mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}(x_1,x_2,x_3)}{\delta a_\mu ^b(x)}}\right)^{[0]}`$ $`=`$ $`0\left|\text{}T\psi (x_1)\psi (x_2)\psi (x_3)A^{b\mu }(x)\right|N_\alpha ^a(\stackrel{}{P}).`$ By using the canonical operator analysis, i.e., by inserting into the Green’s function the complete set relation in the baryon number 1 sector: $$\text{1}\text{ }\text{ }=\underset{a\alpha }{}\frac{d^3p}{(2\pi )^3}\frac{m_N}{E_N(\stackrel{}{p}^2)}|\text{}N_\alpha ^a(\stackrel{}{p})\text{}N_\alpha ^a(\stackrel{}{p})|+\mathrm{},$$ (15) and by using the Fourier transform of the step function $`\theta (t)`$, we see that<sup>6</sup><sup>6</sup>6The argument is similar to the one given in p.92 in . the product $`\mathrm{\Psi }^{[0]}\overline{\mathrm{\Psi }}^{[0]}`$ provides the residue of the Green’s function at the nucleon pole, i.e., $$G^{[0]}=\underset{a\alpha }{}\frac{d^4p}{(2\pi )^4}\frac{m_N}{E_N(\stackrel{}{p}^2)}\frac{i}{p_0E_N(\stackrel{}{p}^2)+iϵ}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{p})]^{[0]}\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{p})]^{[0]}+\mathrm{}$$ (16) In Appendix A, we explain how to obtain and normalize these two 3qBS amplitudes based on the relativistic Faddeev framework. In Appendix A, we prove that the residues $`\mathrm{\Psi }^{[0]}`$ and $`\overline{\mathrm{\Psi }}^{[0]}`$ satisfy the following 3qBS equation in the absence of the external fields(cf Eq.(6)) $$K^{[0]}\mathrm{\Psi }^{[0]}[N_\alpha ^a(\stackrel{}{P})]=\mathrm{\Psi }^{[0]}[N_\alpha ^a(\stackrel{}{P})],\stackrel{~}{\overline{\mathrm{\Psi }}}^{[0]}[N_\alpha ^a(\stackrel{}{P})]K^{[0]}=\stackrel{~}{\overline{\mathrm{\Psi }}}^{[0]}[N_\alpha ^a(\stackrel{}{P})],$$ (17) where a tilde is used to indicate an “amputation”, i.e., $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{p})]^{[0]}=\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{p})]^{[0]}G_0^{[0]1}`$. By rearranging Eq.(13), we obtain the following relation: $$\left(\frac{i\delta \mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}=\frac{1}{1K^{[0]}}\left(\frac{i\delta K^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[0]}.$$ (18) On the one hand, near the nucleon pole the resolvent is given as $`{\displaystyle \frac{1}{1K^{[0]}}}`$ $`=`$ $`G^{[0]}G_0^{[0]1}`$ $``$ $`{\displaystyle \underset{a,\alpha }{}}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{m_N}{E_N(\stackrel{}{p}^2)}\frac{i}{p^0E_N(\stackrel{}{p}^2)+iϵ}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{p})]^{[0]}\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{p})]^{[0]}}.`$ The l.h.s. of Eq.(18), on the other hand, is given near the nucleon pole as follows: $`\left({\displaystyle \frac{i\delta \mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[e]}}{\delta a_\mu ^b(x)}}\right)^{[0]}`$ $``$ $`{\displaystyle \underset{a^{}\alpha ^{}}{}}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{m_N}{E_N(\stackrel{}{p}^2)}\frac{i}{p^0E_N(\stackrel{}{p}^2)+iϵ}0\left|\text{}T\psi \psi \psi \right|N_\alpha ^{}^a^{}(\stackrel{}{p})N_\alpha ^{}^a^{}(\stackrel{}{p})\left|A_\mu ^b(x)\right|N_\alpha ^a(\stackrel{}{P})}.`$ $`=`$ $`{\displaystyle \underset{a^{}\alpha ^{}}{}}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{m_N}{E_N(\stackrel{}{p}^2)}\frac{i}{p^0E_N(\stackrel{}{p}^2)+iϵ}\mathrm{\Psi }[N_\alpha ^{}^a^{}(\stackrel{}{p})]^{[0]}N_\alpha ^{}^a^{}(\stackrel{}{p})\left|A_\mu ^b(x)\right|N_\alpha ^a(\stackrel{}{P})}.`$ By comparing both sides of Eq.(18), we are left with: $$N_\alpha ^{}^a^{}(\stackrel{}{P}^{})\left|A_\mu ^b(x)\right|N_\alpha ^a(\stackrel{}{P})=\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^{}^a^{}(\stackrel{}{P}^{})]^{[0]}\left(\frac{i\delta K^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[0]}.$$ (21) A similar consideration leads to the following relation: $$N_\alpha ^{}^a^{}(\stackrel{}{P}^{})\left|\text{}i\overline{\psi }(x)\gamma _5\frac{\tau ^b}{2}\psi (x)\right|N_\alpha ^a(\stackrel{}{P})=\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^{}^b^{}(\stackrel{}{P}^{})]^{[0]}\left(\frac{i\delta K^{[e]}}{\delta p^b(x)}\right)^{[0]}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[0]}.$$ (22) A combination of these two relations provides us with: $`N^a^{}(\stackrel{}{P}^{})\left|\text{}_\mu A^{b\mu }(x)2im_0\overline{\psi }(x)\gamma _5\tau ^b\psi (x)\right|N^a(\stackrel{}{P})`$ $`=`$ $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^{}^a^{}(\stackrel{}{P}^{})]^{[0]}\left(\left(_\mu {\displaystyle \frac{i\delta }{\delta a_\mu ^b(x)}}2m_0{\displaystyle \frac{i\delta }{\delta p^b(x)}}\right)K^{[e]}\right)^{[0]}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]^{[0]}.`$ These expressions show that the 3qBS amplitudes provide enough informations to evaluate these matrix elements. Here the “relative time” dependence of the 3qBS amplitude surely plays a very important role. At any rate, we should keep in mind that, although bound state matrix elements are obtained by sandwiching one-body operators in the canonical operator formalism, it is these two-body operators shown on the r.h.s. of Eq.(2) that should be sandwiched in the 3qBS framework<sup>7</sup><sup>7</sup>7If 3PI interactions are included in the 3qBS kernel, there appear three-body operators in addition to the two-body operators in the 3qBS framework., which leads to crucial differences. The corresponding formula to evaluate the bound state matrix elements in the Faddeev framework is derived in Appendix A. Note that it is not hard to extend these arguments beyond the NJL model. If the full 3qBS kernel were at our disposal, the chiral symmetry of the original Lagrangian should directly lead to the PCAC relation in the 3qBS framework. However, because the 3qBS kernel is in practice truncated such as to be manageable, it may happen that the truncation scheme spoils the chiral symmetry. We are thus interested in the criterion to decide which 3qBS kernel gives rise to the PCAC relation correctly. For this purpose, it is convenient to introduce the local “axial” gauge transformation of the external fields as follows: $`i\overline{)}\overline{)}v^{(\omega )}(x)\overline{)}a^{(\omega )}(x)\gamma _5m^{(\omega )}`$ $`=`$ $`\mathrm{\Omega }(x)\left(\text{}i\overline{)}\overline{)}v(x)\overline{)}a(x)\gamma _5m\right)\mathrm{\Omega }(x),`$ (24) $`\mathrm{\Omega }(x)`$ $`=`$ $`e^{i\gamma _5\omega (x)};\omega (x)=\omega ^a(x){\displaystyle \frac{\tau ^a}{2}},`$ where $`v^{(\omega )}`$, $`a^{(\omega )}`$ and $`m^{(\omega )}`$ are the gauge images of $`v`$, $`a`$ and $`m`$, respectively. We assume that $`\omega (x)`$ is also localized in space-time. Infinitesimally, these gauge transformations are expressed as follows: $`{\displaystyle \frac{\delta v_\mu ^a(y)}{\delta \omega ^c(x)}}`$ $`=`$ $`ϵ_{abc}a_\mu ^b(y)\delta ^{(4)}(xy)`$ (25) $`{\displaystyle \frac{\delta a_\mu ^a(y)}{\delta \omega ^c(x)}}`$ $`=`$ $`\delta _{ac}{\displaystyle \frac{}{x^\mu }}\delta ^{(4)}(xy)+ϵ_{abc}v_\mu ^b(y)\delta ^{(4)}(xy)`$ $`{\displaystyle \frac{\delta s(y)}{\delta \omega ^c(x)}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}p^c(y)\delta ^{(4)}(xy)`$ $`{\displaystyle \frac{\delta p^b(y)}{\delta \omega ^c(x)}}`$ $`=`$ $`2\delta _{bc}s(y)\delta ^{(4)}(xy),`$ where $`ϵ_{abc}`$ is the totally anti-symmetric tensor. Let $`F[v,a,m]`$ be a functional. We define the functional derivative $`{\displaystyle \frac{\delta }{\delta \omega ^c(x)}}`$ through the following relation: $$\frac{\delta F}{\delta \omega ^c(x)}\frac{\delta F[v^{(\omega )},a^{(\omega )},m^{(\omega )}]}{\delta \omega ^c(x)}.$$ (26) By using the chain rule, $`{\displaystyle \frac{i\delta }{\delta \omega ^c(x)}}`$ is expressed as a linear combination of functional derivatives with respect to the external fields: $`{\displaystyle \frac{i\delta }{\delta \omega ^c(x)}}`$ $`=`$ $`{\displaystyle d^4y\left[\frac{\delta a_\mu ^b(y)}{\delta \omega ^c(x)}\frac{i\delta }{\delta a_\mu ^b(y)}+\frac{\delta v_\mu ^b(y)}{\delta \omega ^c(x)}\frac{i\delta }{\delta v_\mu ^b(y)}+\frac{\delta s(y)}{\delta \omega ^c(x)}\frac{i\delta }{\delta s(y)}+\frac{\delta p^b(y)}{\delta \omega ^c(x)}\frac{i\delta }{\delta p^b(y)}\right]}`$ $`=`$ $`_\mu {\displaystyle \frac{i\delta }{\delta a_\mu ^c(x)}}ϵ_{abc}v_\mu ^a(x){\displaystyle \frac{i\delta }{\delta a_\mu ^b(x)}}ϵ_{abc}a_\mu ^a(x){\displaystyle \frac{i\delta }{\delta v_\mu ^b(x)}}+{\displaystyle \frac{1}{2}}p^c(x){\displaystyle \frac{i\delta }{\delta s(x)}}2s(x){\displaystyle \frac{i\delta }{\delta p^c(x)}}.`$ In particular, we have the following relation in the vanishing external field limit: $$\left(\frac{i\delta K^{[e^{(\omega )}]}}{\delta \omega ^b(x)}\right)^{[0]}=_\mu \left(\frac{i\delta K^{[e]}}{\delta a_\mu ^b(x)}\right)^{[0]}2m_0\left(\frac{i\delta K^{[e]}}{\delta p^b(x)}\right)^{[0]},$$ (28) where $`[e^{(\omega )}]`$ denotes the gauge transformed fields $`[v^{(\omega )},a^{(\omega )},m^{(\omega )}]`$. Now we state the sufficient condition for the validity of the PCAC relation as follows: Sufficient Condition: If a 3qBS kernel behaves in the following manner under a gauge transformation $`\mathrm{\Omega }(x)`$: $`K^{[e^{(\omega )}]}(x_1,x_2,x_3;y_1,y_2,y_3)`$ $`=`$ $`\left(\text{}\mathrm{\Omega }(x_1)^1\mathrm{\Omega }(x_2)^1\mathrm{\Omega }(x_3)^1\right)K^{[e]}\left(\text{}\mathrm{\Omega }(y_1)\mathrm{\Omega }(y_2)\mathrm{\Omega }(y_3)\right),`$ then the PCAC relation is satisfied. Our “criterion” is obtained from Eq.(2) by applying the functional derivative with respect to the gauge transformation on both sides, and using Eq.(28). We are left with the following chiral Ward identity, which is the “criterion” for the validity of the PCAC relation: Criterion: (The Chiral Ward Identity for the 3qBS Kernel) $`\left[\left(_\mu {\displaystyle \frac{i\delta }{\delta a_\mu ^b(x)}}2m_0{\displaystyle \frac{i\delta }{\delta p^b(x)}}\right)K^{[e]}\right]^{[0]}(x_1,x_2,x_3;y_1,y_2,y_3)`$ (30) $`=`$ $`\begin{array}{c}\text{}K^{[0]}\text{}\left[\text{}(\tau ^b\gamma _511)\delta ^{(4)}(xy_1)+(1\tau ^b\gamma _51)\delta ^{(4)}(xy_2)+(11\tau ^b\gamma _5)\delta ^{(4)}(xy_3)\right]\hfill \\ \text{}\left[(\tau ^b\gamma _511)\delta ^{(4)}(xx_1)+(1\tau ^b\gamma _51)\delta ^{(4)}(xx_2)+(11\tau ^b\gamma _5)\delta ^{(4)}(xx_3)\right]K^{[0]}.\hfill \end{array}`$ (33) To see that this indeed leads to the PCAC relation, we sandwich Eq.(30) between $`\mathrm{\Psi }[N_\alpha ^{}^a^{}(\stackrel{}{P}^{})]^{[0]}`$ and $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{P})]^{[0]}`$ and use Eq.(17) and Eq.(2) to get $$N_\alpha ^{}^a^{}(\stackrel{}{P}^{})\left|\text{}_\mu A^{b\mu }(x)2m_0i\overline{\psi }(x)\gamma _5\tau ^b\psi (x)\right|N_\alpha ^a(\stackrel{}{P})=0,$$ (34) which is the desired relation. We should comment here on the reason why we called Eq.(2) a “sufficient condition”. If we try to extend our method to $`SU(3)_L\times SU(3)_R`$-QCD in the presence of external chiral gauge fields, the relation (2) is no longer valid due to the existence of the non-Abelian anomaly<sup>8</sup><sup>8</sup>8This anomaly should not be confused with the $`U_A(1)`$ anomaly, i.e., the Abelian anomaly. . Note, however, that, even if there is such a non-Abelian anomaly, because the anomalous contributions are polynomials of at least second order in the external chiral gauge fields, the infinitesimal form (the criterion Eq.(30)) is still valid. However, if more than two functional derivatives of the external chiral gauge fields are involved, one must pay full attention to the non-Abelian anomaly even for the infinitesimal form. At any rate, the chiral $`SU(2)_L\times SU(2)_R`$ case, which is of most interest to us here, is known to be anomaly free. ## 3 Feynman Diagrams The aim of this section is to see which kinds of Feynman diagrams are involved in the expression Eq.(21) of $`N|A_\mu ^b(x)|N`$, before applying the criterion/sufficient condition to particular 3qBS kernels. We first consider the constituent quark propagator $`S_F^{[e]}(x,y)`$ in the presence of the external fields. The self energy is, in principle, obtained from the sum of 1PI diagrams. However, in practice, it is approximated in the mean field (Hartree-Fock) treatment, as expressed by the following gap equation: $`iS_F^{[e]}(x,y)`$ $`=`$ $`iS_{0;F}^{[e]}(x,y)`$ $``$ $`{\displaystyle \underset{\mathrm{\Gamma }}{}}2ig_\mathrm{\Gamma }^{(q\overline{q})}{\displaystyle d^4ziS_{0;F}^{[e]}(x,z)\mathrm{\Gamma }iS_F^{[e]}(z,y)\text{Tr}\left(\mathrm{\Gamma }iS_F^{[e]}(z,z)\right)}.`$ $`g_\mathrm{\Gamma }^{(q\overline{q})}`$ are the effective coupling constants in the $`q\overline{q}`$ channels, which are defined in Appendix B.1. $`S_{0;F}^{[e]}(x,y)`$ is the current quark propagator, which is defined through the following relation: $$\left(i\overline{)}^{(x)}\overline{)}a(x)\gamma _5\overline{)}v(x)m(x)\right)S_{0;F}^{[e]}(x,y)=\delta ^{(4)}(xy).$$ (36) The solution to Eq.(3) is obtained as a self-consistent solution to the following equations: $`S_F^{[e]}(x,y)`$ $`=`$ $`\left(i\overline{)}\overline{)}a\gamma _5\overline{)}vm\mathrm{\Sigma }^{[e]}\right)^1(x,y)`$ (37) $`\mathrm{\Sigma }^{[e]}(z)`$ $``$ $`{\displaystyle \underset{\mathrm{\Gamma }}{}}2ig_\mathrm{\Gamma }^{(q\overline{q})}\text{Tr}\left(S_F^{[e]}(z,z)\mathrm{\Gamma }\right)\mathrm{\Gamma }.`$ $`\mathrm{\Sigma }^{[e]}(z)`$ is the self-energy of the constituent quark. Now, several comments are in order. (1) Due to the presence of the external fields, there is no translational symmetry any more, and therefore the self-energy depends on the space-time point $`z`$. (The dependence on just a single space-time coordinate $`z`$ is due to the mean field approximation.) (2) Non-vanishing external fields may lead to not only a non-vanishing scalar condensate, but also a non-vanishing pseudoscalar, vector, axial vector condensates. However, in the vanishing external field limit, we have $`m_0+\mathrm{\Sigma }^{[0]}=M`$ (the constituent quark mass). (3) Because the NJL model is non-renormalizable, it is necessary to introduce a UV regularization in all loop integrals. Hereafter, whenever a loop integral is encountered, the integration is understood as a regularized one. (4) If this regularization respects the chiral symmetry, $`S_F^{[e]}(x,y)`$ and $`\mathrm{\Sigma }^{[e]}(z)`$ should behave in the following way: $`S_F^{[e^{(\omega )}]}(x,y)`$ $`=`$ $`\mathrm{\Omega }(x)^1S_F^{[e]}(x,y)\mathrm{\Omega }(y)^1`$ (38) $`\mathrm{\Sigma }^{[e^{(\omega )}]}(z)`$ $`=`$ $`\mathrm{\Omega }(z)\mathrm{\Sigma }^{[e]}(z)\mathrm{\Omega }(z).`$ For later convenience, the transformation of $`S_F`$ is depicted in Fig.1. All diagrammatically truncated 3qBS kernels $`K^{[e]}`$ can be expressed as the product of several constituent quark propagators $`S_F^{[e]}`$ and several elementary local vertices $`\mathrm{\Gamma }`$. Because, in our case, these elementary local vertices do not depend on the external fields<sup>9</sup><sup>9</sup>9If derivative couplings are involved in the interaction Lagrangian, our argument is straightforwardly extended by replacing ordinary derivatives by chiral covariant derivatives. In this case, $`\delta /\delta a_\mu ^b(x)`$ also hits the argument $`\mathrm{\Gamma }`$ in $`K[\mathrm{\Gamma }^{[e]};S_F^{[e]}]`$. This is essential to extend the arguments given in the next section. In this case, it is not simply $`A^{b\mu }(x)`$ but the conserved current operator, that is inserted in the r.h.s. in the 3rd line in Eq.(14) — $`A^{b\mu }(x)`$ fails to be a conserved current in this case., $`K^{[e]}`$ depends on $`v,a`$ and $`m`$ only through $`S_F^{[e]}`$. In a symbolic notation, this may be denoted as $`K^{[e]}=K[\mathrm{\Gamma };S_F^{[e]}]`$, and the functional derivatives are then symbolically expressed by using the chain rule as follows: $$\left[\frac{i\delta K^{[e]}}{\delta a_\mu ^b(x)}\right]^{[0]}=\underset{\alpha \alpha ^{}}{}d^4zd^4z^{}\left[\frac{i\delta S_{F}^{}{}_{\alpha \alpha ^{}}{}^{[e]}(z,z^{})}{\delta a_\mu ^b(x)}\right]^{[0]}\left(\frac{\delta K[\mathrm{\Gamma };S_F^{[0]}]}{\delta S_{F}^{[0]}{}_{\alpha \alpha ^{}}{}^{}(z,z^{})}\right),$$ (39) where $`\alpha ,\alpha ^{}`$ are triples of the Dirac, iso-spin, and color indices. This relation implies that the general rule to obtain the functional derivative of the 3qBS kernel is to replace each constituent quark propagator in turn by its derivative, and then sum up the resulting terms. Now all we need is $`\delta S_F^{[e]}/\delta a_\mu ^b(x)`$, which is obtained by applying $`\delta /\delta a_\mu ^b(x)`$ to Eq.(37) as follows: $`\left[{\displaystyle \frac{\delta S_F^{[e]}(x,y)}{\delta a_\mu ^b(z)}}\right]^{[0]}`$ $`=`$ $`\begin{array}{c}S_F^{[0]}(x,z)\left(\text{}\gamma ^\mu \gamma _5{\displaystyle \frac{\tau ^b}{2}}\right)S_F^{[0]}(z,y)\hfill \\ +{\displaystyle d^4z^{}S_F^{[0]}(x,z^{})\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta a_\mu ^b(z)}\right]^{[0]}S_F^{[0]}(z^{},y)}.\hfill \end{array}`$ (42) $`\left[{\displaystyle \frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta a_\mu ^b(z)}}\right]^{[0]}`$ $`=`$ $`{\displaystyle \underset{\mathrm{\Gamma }}{}}2ig_\mathrm{\Gamma }^{(q\overline{q})}\text{Tr}\left(\left[{\displaystyle \frac{\delta S_F^{[e]}(z^{},z^{})}{\delta a_\mu ^b(z)}}\right]^{[0]}\mathrm{\Gamma }\right)\mathrm{\Gamma }.`$ (43) By inserting the first relation into the second, we obtain the following closed equation for $`\delta \mathrm{\Sigma }/\delta a_\mu ^b`$: $$\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta a_\mu ^b(z)}\right]^{[0]}=\begin{array}{c}\underset{\mathrm{\Gamma }}{}2ig_\mathrm{\Gamma }^{(q\overline{q})}\text{Tr}\left(S^{[0]}(z^{},z)\left(\text{}\gamma ^\mu \gamma _5\frac{\tau ^b}{2}\right)S^{[0]}(z,z^{})\mathrm{\Gamma }\right)\mathrm{\Gamma }\hfill \\ \text{}+d^4z^{\prime \prime }\underset{\mathrm{\Gamma }}{}2ig_\mathrm{\Gamma }^{(q\overline{q})}\text{Tr}\left(S^{[0]}(z^{},z^{\prime \prime })\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{\prime \prime })}{\delta a_\mu ^b(z)}\right]^{[0]}S^{[0]}(z^{\prime \prime },z^{})\mathrm{\Gamma }\right)\mathrm{\Gamma }.\hfill \end{array}$$ (44) For simplicity, we adopt<sup>10</sup><sup>10</sup>10The non-vanishing effective $`q\overline{q}`$ coupling constant in the iso-vector axial-vector channel $`g_{\mathrm{ax}}`$ does not change the chiral symmetry properties. the effective $`q\overline{q}`$ coupling constant in the iso-vector axialvector channel $`g_{\mathrm{ax}}=0`$. (For the precise meaning of $`g_{\mathrm{ax}}`$, see Appendix B. We will discuss the general case, i.e., $`g_{\mathrm{ax}}0`$, in Appendix D.) We can parameterize our solution as follows: $$d^4z^{}e^{iq(z^{}z)}\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta a_\mu ^b(z)}\right]^{[0]}=\stackrel{~}{H}(q^2)\left(q_\mu \gamma _5\frac{\tau ^b}{2}\right),$$ (45) and from Eq.(44) we see that $`\stackrel{~}{H}(q^2)`$ satisfies the following equation: $$\stackrel{~}{H}(q^2)=2ig_\pi \mathrm{\Pi }_{5A}(q^2)2ig_\pi \mathrm{\Pi }_{55}(q^2)\stackrel{~}{H}(q^2),$$ (46) where $`\mathrm{\Pi }_{5A}(q^2)`$ and $`\mathrm{\Pi }_{55}(q^2)`$ are the bubble integrals which are defined in Appendix B.2. The solution is expressed as a geometric series: $$\stackrel{~}{H}(q^2)=\frac{2ig_\pi \mathrm{\Pi }_{5A}(q^2)}{1+2ig_\pi \mathrm{\Pi }_{55}(q^2)}.$$ (47) $`\delta S_F/\delta a_\mu ^b(z)`$ is depicted in Fig.2.(a). The two diagrams on the r.h.s. correspond to the two terms in Eq.(42), respectively, where the second one is proportional to $`q^\mu \stackrel{~}{H}(q^2)`$ and has the pionic pole. Note that, whenever the quark propagator is obtained as a non-trivial solution to a self-consistent equation which leads to the chiral symmetry breaking, $`\delta S_F/\delta a_\mu ^b`$ always has a non-trivial “mesonic part” $`\stackrel{~}{H}(q^2)`$. Now we can consider particular 3qBS kernels and the associated diagrams relevant for the matrix element calculation. Our first example is the ladder truncated 3qBS kernel, which is depicted in Fig.3(a). Solid lines represent the constituent quark propagator. A slash indicates amputation of the constituent quark propagator. The interaction strengths in the various $`qq`$ channels are obtained by applying the Fierz identity to the interaction Lagrangian. (See Appendix B.) In this case, Eq.(39) leads to the diagram in Fig.3(b), where the insertion on the quark line indicated by the “$``$” has been defined in Fig.2. (Diagrams with the same topologies are omitted.) The second example is the $`qq`$ interaction involving the exchange of $`q\overline{q}`$ pairs in the t-channel (“meson exchange interaction”), which is depicted in Fig.4(a). In this case, Eq.(39) leads to the diagrams depicted in Fig.4(b). We see that, in addition to the coupling of the external field to the external quark propagators, couplings to internal quark propagators are also involved (the second diagram), which are often referred to as the “meson exchange current” contributions. It is straightforward to extend these considerations to more complicated and realistic cases. A moral is that, once the 3qBS kernel is specified, a unique set of Feynman diagrams exists to determine the matrix element of the axial vector current. ## 4 Applications of the Criterion/Sufficient Condition The aims of this section are to see whether the 3qBS kernels presented in the previous section satisfy the criterion Eq.(30)/sufficient condition Eq.(2), and to provide examples how to use the general considerations of Section 2 practically. We first consider the chiral transformation properties of the elementary local vertices in the Lagrangian. The global chiral symmetry of the interaction Lagrangian $`_I={\displaystyle \underset{\mathrm{\Gamma }}{}}g_\mathrm{\Gamma }(\overline{\psi }\mathrm{\Gamma }\psi )^2`$ implies the following identity: $$\underset{\mathrm{\Gamma }}{}g_\mathrm{\Gamma }\left(e^{i\gamma _5\tau ^b\mathrm{\Theta }^b}\mathrm{\Gamma }e^{i\gamma _5\tau ^b\mathrm{\Theta }^b}\right)_{ij}\left(e^{i\gamma _5\tau ^b\mathrm{\Theta }^b}\mathrm{\Gamma }e^{i\gamma _5\tau ^b\mathrm{\Theta }^b}\right)_{kl}=\underset{\mathrm{\Gamma }}{}g_\mathrm{\Gamma }\mathrm{\Gamma }_{ij}\mathrm{\Gamma }_{kl}.$$ (48) Since $`_I`$ is a contact interaction without any derivative terms, this identity remains valid even if $`\mathrm{\Theta }^b`$ acquires space–time dependence, i.e., $$\underset{\mathrm{\Gamma }}{}g_\mathrm{\Gamma }\left(\text{}\mathrm{\Omega }(x)\mathrm{\Gamma }\mathrm{\Omega }(x)\right)_{ij}\left(\text{}\mathrm{\Omega }(x)\mathrm{\Gamma }\mathrm{\Omega }(x)\right)_{kl}=\underset{\mathrm{\Gamma }}{}g_\mathrm{\Gamma }\mathrm{\Gamma }_{ij}\mathrm{\Gamma }_{kl}.$$ (49) Actually, these $`q\overline{q}`$ interactions consist of several “closed chiral multiplet sectors”, and the identity of the type Eq.(49) holds in each such sector separately. For example, the $`q\overline{q}`$ interaction in the $`\pi `$ and $`\sigma `$ mesonic channel forms a closed chiral multiplet under the local axial gauge transformation, i.e., the following identity holds: $`\delta _{ij}\delta _{kl}{\displaystyle \underset{a=1}{\overset{3}{}}}\left(\gamma _5\tau ^a\right)_{ij}\left(\gamma _5\tau ^a\right)_{kl}`$ $`=`$ $`\begin{array}{c}\left(\text{}\mathrm{\Omega }(x)\mathrm{\Omega }(x)\right)_{ij}\left(\text{}\mathrm{\Omega }(x)\mathrm{\Omega }(x)\right)_{kl}\hfill \\ \text{}{\displaystyle \underset{a=1}{\overset{3}{}}}\left(\text{}\mathrm{\Omega }(x)\gamma _5\tau ^a\mathrm{\Omega }(x)\right)_{ij}\left(\text{}\mathrm{\Omega }(x)\gamma _5\tau ^a\mathrm{\Omega }(x)\right)_{kl}.\hfill \end{array}`$ (52) We need to establish similar relations in the $`qq`$ sector. For this purpose, it is convenient to use the following Fierz identity (for details, see Appendix B.1.): $$_I=\underset{\mathrm{\Gamma }}{}g_\mathrm{\Gamma }(\overline{\psi }\mathrm{\Gamma }\psi )^2=\underset{\mathrm{\Gamma }^{}}{}g_\mathrm{\Gamma }^{}^{(qq)}(\overline{\psi }\mathrm{\Gamma }^{}\overline{\psi }^T)(\psi ^T\mathrm{\Gamma }^{}\psi ).$$ (53) In this representation, the identity Eq.(49) is expressed in the following way: $$\underset{\mathrm{\Gamma }^{}}{}g_\mathrm{\Gamma }^{}^{(qq)}\left(\mathrm{\Omega }(x)\mathrm{\Gamma }^{}\mathrm{\Omega }(x)^T\right)_{ij}\left(\mathrm{\Omega }(x)^T\mathrm{\Gamma }^{}\mathrm{\Omega }(x)\right)_{kl}=\underset{\mathrm{\Gamma }^{}}{}g_\mathrm{\Gamma }^{}^{(qq)}\mathrm{\Gamma }_{ij}^{}\mathrm{\Gamma }_{kl}^{}.$$ (54) These $`qq`$ interactions also consist of several closed chiral multiplet sectors, and in each sector Eq.(54) is valid separately. In particular, the $`qq`$ interaction in the scalar diquark channel ($`J^\pi =0^+`$, isoscalar, color $`\overline{3}`$) forms a closed chiral singlet, i.e., $`{\displaystyle \underset{A=2,5,7}{}}\left(\text{}(\gamma _5C^1)\tau _2\beta _A\right)_{ij}\left(\text{}(C\gamma _5)\tau _2\beta _A\right)_{kl}`$ $`=`$ $`{\displaystyle \underset{A=2,5,7}{}}\left(\text{}\mathrm{\Omega }(x)\left(\text{}(\gamma _5C^1)\tau _2\beta _A\right)\mathrm{\Omega }(x)^T\right)_{ij}\left(\text{}\mathrm{\Omega }(x)^T\left(\text{}(C\gamma _5)\tau _2\beta _A\right)\mathrm{\Omega }(x)\right)_{kl},`$ where $`\beta ^A`$ is the rescaled color Gell-Mann matrix $`\beta _A=\sqrt{\frac{3}{2}}\lambda _A`$ with the normalization $`\text{tr}(\beta _A\beta _B)=3\delta _{AB}`$. Note that $`\beta _A`$ for $`A=2,5,7`$ are anti-symmetric matrices corresponding to the color $`\overline{3}_c`$ diquark channels. The $`qq`$ interaction in the axial vector diquark ($`J^\pi =1^+`$, isovector, color $`\overline{3}`$) together with the vector diquark ($`J^\pi =1^{}`$, isoscalar, color $`\overline{3}`$) channel forms a closed chiral multiplet, i.e., $`{\displaystyle \underset{A=2,5,7}{}}\begin{array}{c}[{\displaystyle \underset{a=1}{\overset{3}{}}}\left(\text{}(\gamma _\mu C^1)(\tau _a\tau _2)\beta _A\right)_{ij}\left(\text{}(C\gamma ^\mu )(\tau _2\tau _a)\beta _A\right)_{kl}\hfill \\ \text{}\left(\text{}(\gamma _\mu \gamma _5C^1)\tau _2\beta _A\right)_{ij}\left(\text{}(C\gamma _5\gamma ^\mu )\tau _2\beta _A\right)]\hfill \end{array}`$ (58) $`=`$ $`{\displaystyle \underset{A=2,5,7}{}}\begin{array}{c}[{\displaystyle \underset{a=1}{\overset{3}{}}}\left(\text{}\mathrm{\Omega }(x)\left(\text{}(\gamma _\mu C^1)(\tau _a\tau _2)\beta _A\right)\mathrm{\Omega }(x)^T\right)_{ij}\left(\text{}\mathrm{\Omega }(x)^T\left(\text{}(C\gamma ^\mu )(\tau _2\tau _a)\beta _A\right)\mathrm{\Omega }(x)\right)_{kl}\hfill \\ \text{}\left(\text{}\mathrm{\Omega }(x)\left(\text{}(\gamma _\mu \gamma _5C^1)\tau _2\beta _A\right)\mathrm{\Omega }(x)^T\right)_{ij}\left(\text{}\mathrm{\Omega }(x)^T\left(\text{}(C\gamma _5\gamma ^\mu )\tau _2\beta _A\right)\mathrm{\Omega }(x)\right)_{kl}]\hfill \end{array}`$ (61) We give a list of these closed chiral multiplets in Appendix B.1. Now we consider whether the sufficient condition Eq.(2) is satisfied in the case of the ladder truncation scheme, i.e., the 3qBS kernel of the type Fig.3. In this case, the 2PI interaction consists of only an elementary contact interaction. In practical numerical calculations, it is further truncated according to the quantum numbers of the diquark channels, i.e., the scalar diquark channel, the axialvector diquark channel, etc. We assume that the truncated vertex corresponds to a sum of closed chiral multiplets. Fig.5 shows the steps involved in the analysis of the gauge transformation properties of the kernel. In the first step, we apply the local gauge transformation $`\mathrm{\Omega }(x)`$ in order to get the transformed kernel (l.h.s.of Eq.(2)). Only the constituent quark propagators transform, and their transformation is given by Eq.(38). In the second step, we use Eq.(54), and the last step is due to the fact that amputated propagators are delta functions. As a result, the condition Eq.(2) is satisfied by the kernel in the ladder approximation, provided that the truncation of $`qq`$ channels is done such as to have closed chiral multiplets. In particular, since the $`qq`$ interaction in the scalar diquark channel forms a chiral singlet, the 3qBS framework in the ladder truncation scheme keeping only the $`qq`$ interaction in the scalar diquark channel gives rise the PCAC relation correctly. However, if the $`qq`$ interaction in the axialvector diquark channel is further included, it is in principle necessary to include also the $`qq`$ interaction in the vector diquark channel. In practice, however, the vector diquark channel can be expected to have small effects, at least on the nucleon mass, from the nonrelativistic analogy. Next we consider whether the sufficient condition is satisfied in the case of the kernel involving $`q\overline{q}`$ exchange, i.e., the 3qBS kernel of the type Fig.4. Fig.6 shows the argument for one of the infinite terms involved in the ladder sum. We assume that each vertex corresponds to a sum of closed chiral multiplets. In the first step we apply the gauge transformation using Eq.(38) to transform the constituent quark propagator. In the second step we use Eq.(49). Note that all the phase factors, which appear around the internal vertices, cancel themselves. The last step is due to the fact that the amputated propagator is just a delta function. This demonstrates the validity of Eq.(2) for the kernel involving $`q\overline{q}`$ exchange. Several comments are in order. First, since the mass of the pion is so small, the $`q\overline{q}`$ exchange in the pionic channel will contribute much more than the one in the sigma mesonic channel. However, in principle, it is necessary to include both channels in order to get the PCAC relation correctly. Second, if we include the $`q\overline{q}`$ exchange interaction in the 3qBS kernel, it is necessary to take into account also the “meson exchange current contributions” in the calculation of the matrix element of the axial current, since the l.h.s. of our criterion Eq.(30) involves also the second diagram in Fig.4.(b), as discussed in Section 3. These examples are straightforwardly extended to more general, complicated and realistic cases (for example, to the expansion scheme of ). We may consider these 3qBS kernel as a formal sum of 2PI and 3PI diagrams. It is thus not hard to convince ourselves that the chiral symmetry in the original interaction Lagrangian leads to the PCAC relation correctly in our formalism, provided the truncation of the kernel is done consistent with the condition (2). ## 5 The Goldberger-Treiman Relation and $`g_{\pi NN}`$ In this section, we give the explicit numerical results for $`g_{\pi NN}`$ in the chiral limit together with the off-shell $`g_{\pi NN}`$ off the chiral limit. We restrict our attention to the ladder truncation scheme keeping only the $`qq`$ interaction in the scalar diquark channel. Although this truncation scheme of 3qBS kernel should lead to the correct Goldberger-Treiman relation, the UV regularization (the Euclidean cutoff in our case) unfortunately violates the chiral symmetry. One of the aims of this section is therefore to estimate the degree of violation of the PCAC/Goldberger-Treiman relation due to the UV-regularization. In this section, unless explicitly indicated for the more general case, the external fields are understood to be absent, i.e., $`v=a=0`$, $`m=m_0`$. For consistency reason, we prefer to use the axial current operator $`A_\mu ^b(x)`$ as an interpolating field for the pion in this paper. For those readers who prefer to use $`\overline{\psi }(x)\gamma _5{\displaystyle \frac{\tau ^b}{2}}\psi (x)`$ as an interpolating field for the pion, the following relation would be convenient<sup>11</sup><sup>11</sup>11We will not use this relation explicitly in this paper.: $$A_\mu ^b(x)=f_\pi _\mu \pi _{\mathrm{as}}^b(x)+\mathrm{},$$ (62) where $`\pi _{\mathrm{as}}^b(x)`$ is the asymptotic field operator for the pion. ### 5.1 The Extraction of $`g_{\pi NN}`$ in the Chiral Limit In this subsection, $`m_0=0`$ is understood. The axial form factors $`g_A(q^2)`$ and $`h_A(q^2)`$ are defined through the following relation: $$\overline{u}^{(\alpha )a}(\stackrel{}{P})\frac{\tau ^b}{2}\left(\text{}g_A(q^2)\gamma ^\mu \gamma _5+q^\mu h_A(q^2)\gamma _5\right)u^{(\beta )c}(\stackrel{}{L})=N^{(\alpha )a}(\stackrel{}{P})\left|\text{}A^{b\mu }(x=0)\right|N^{(\beta )c}(\stackrel{}{L}),$$ (63) where $`\overline{u}^{(\alpha )a}(\stackrel{}{P})`$ and $`\overline{u}^{(\beta )c}(\stackrel{}{L})`$ are the final and initial nucleon bispinors with the isospins $`a,c`$, the helicities $`\alpha ,\beta `$ and the momenta $`P,L`$ ($`P^2=L^2=m_N^2`$), respectively. $`q=PL`$ is the momentum transfer. $`g_A`$ and $`g_{\pi NN}`$ are extracted as follows: $$g_A=g_A(q^2=0),h_A(q^2)=\frac{2f_\pi g_{\pi NN}}{q^2}+\mathrm{}.$$ (64) A multiplication of $`q_\mu `$ on both sides of Eq.(63) leads to $`0`$ $`=`$ $`\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\left(\text{}g_A(q^2)\overline{)}q\gamma _5+q^2h_A(q^2)\gamma _5\right)u^{(\beta )c}(\stackrel{}{L})`$ $`=`$ $`\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\left(\text{}2m_Ng_A(q^2)\gamma _5+q^2h_A(q^2)\gamma _5\right)u^{(\beta )c}(\stackrel{}{L})`$ $`0`$ $`=`$ $`2m_Ng_A(q^2)+q^2h_A(q^2),`$ (66) where we used the PCAC relation (34) for $`m_0=0`$. As is well-known, due to the pion pole of $`h_A(q^2)`$ at $`q^2=0`$, this relation leads to the Goldberger–Treiman(GT) relation: $$m_Ng_A=f_\pi g_{\pi NN}.$$ (67) We expand the both sides of Eq.(63) with respect to a small $`q^2`$ to extract $`g_{\pi NN}`$. The l.h.s. is expanded as follows: $`\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\left(\text{}g_A(q^2)\gamma ^\mu \gamma _5+q^\mu h_A(q^2)\gamma _5\right)u^{(\beta )c}(\mathrm{\Lambda }^1\stackrel{}{P})`$ $`=`$ $`\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\left(\text{}g_A(q^2)\gamma ^\mu \gamma _5+q^\mu h_A(q^2)\gamma _5\right)\widehat{S}(\mathrm{\Lambda }^1)u^{(\beta )c}(\stackrel{}{P})`$ $``$ $`\left(\text{}\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\gamma ^\mu \gamma _5u^{(\beta )b}(\stackrel{}{P})\right)g_A+\left(\text{}\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\gamma _5S(\lambda )u^{(\beta )c}(\stackrel{}{P})\right)q^\mu {\displaystyle \frac{2f_\pi g_{\pi NN}}{q^2}}+O(q),`$ where $`(\mathrm{\Lambda }^1)_{\mu }^{}{}_{}{}^{\nu }=(e^\lambda )_{\mu }^{}{}_{}{}^{\nu }`$ is a boost matrix, i.e., $`L_\mu =(\mathrm{\Lambda }^1)_{\mu }^{}{}_{}{}^{\nu }P_\nu `$, $`\widehat{S}(\mathrm{\Lambda }^1)=e^{S(\lambda )}`$ is the boost matrix in the Dirac bispinor space with $`S(\lambda )=(\lambda )_{\mu \nu }{\displaystyle \frac{i}{4}}\sigma ^{\mu \nu }`$. Note that $`q_\mu =P_\mu (\mathrm{\Lambda }^1)_{\mu }^{}{}_{}{}^{\nu }P^\nu \lambda _{\mu }^{}{}_{}{}^{\nu }P_\nu +O(\lambda ^2)`$. We thus have $`\lambda =O(q)`$. Hence we have $$\left(\text{}\overline{u}^{(\alpha )a}(\stackrel{}{P})\frac{\tau ^b}{2}\gamma _5S(\lambda )u^{(\beta )c}(\stackrel{}{P})\right)q^\mu =O(q^2),$$ (69) which cancels the pion pole of $`h_A(q^2)`$ at $`q^2=0`$. The r.h.s. of Eq.(63) is evaluated by the 3qBS /Faddeev expression Eq.(21) and Eq.(110) of the nucleon matrix element: $`N^{(\alpha )a}(\stackrel{}{P})\left|\text{}A^{b\mu }(x=0)\right|N^{(\beta )c}(\stackrel{}{L})`$ $`=`$ $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{P})]\left[{\displaystyle \frac{i\delta K^{[e]}}{\delta a_\mu ^b(x=0)}}\right]^{[0]}\mathrm{\Psi }[N_\beta ^c(\stackrel{}{L})]`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]\left[{\displaystyle \frac{i\delta K_F^{[e]}}{\delta a_\mu ^b(x=0)}}\right]^{[0]}\psi [N_\beta ^c(\stackrel{}{L})],`$ where $`\psi [N_\beta ^c(\stackrel{}{L})]`$, $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]`$ and $`K_F^{[e]}`$ are the Faddeev amplitudes and the Faddeev kernel, respectively. These quantities are defined in Appendix A, where also their relation to the 3qBS quantities is explained. The Faddeev equation is depicted in Fig.7. Since we adopted the ladder truncation scheme in this section, the $`qq`$ interaction in Fig.7 is understood to be point-like and separable. Hence the double line in Fig.7 can be expressed as a geometric series of $`qq`$ bubble integrals. In addition, since the $`qq`$ interaction is truncated to the scalar diquark channel, the double line in Fig.7 now stands for the following t-matrix $`t_{\mathrm{sd}}(q^2)`$ in the scalar diquark channel: $$t_{\mathrm{sd}}(q^2)=2\times \frac{2ig_{\mathrm{sd}}}{12ig_{\mathrm{sd}}\mathrm{\Pi }_{\mathrm{sd}}(q^2)},$$ (71) with $$\mathrm{\Pi }_{\mathrm{sd}}(q^2)\delta _{AB}=\frac{d^4k}{(2\pi )^4}\text{tr}\left(\text{}(C\gamma _5\tau _2\beta _A)iS_F(q+k)(\gamma _5C^1\tau _2\beta _B)iS_F(k)^T\right).$$ (72) Since the second line in Eq.(5.1) involves the functional derivative of the Faddeev kernel, the contributions are classified as the following three types<sup>12</sup><sup>12</sup>12In the diagrams of Fig.8, we adopted the relative momenta of the spectator quark and the diquark, which are defined in Eq.(143), with the value $`\eta =1/2`$.(cf. Appendix A.3): (a) the quark current contribution\[Fig.8.(a)\], (b) the exchange current contribution\[Fig.8.(b)\], (c) the diquark current contribution\[Fig.8.(c)\]. Note that, due to the iso-scalar nature of the scalar diquark, the diquark current contribution to the matrix element of the iso-vector axial current vanishes identically. The evaluation of the diagram Fig.8.(a) posses a problem, since there appears a delta function $`\delta ^{(4)}(plq/2)`$ in addition to the delta function associated with the total momentum conservation. Therefore, depending on the momentum transfer $`q`$, the evaluation involves the values of the Faddeev amplitude at points which are outside the mesh used for the solution of the Faddeev equation<sup>13</sup><sup>13</sup>13The numerical procedure to obtain the mass and the associated Faddeev amplitudes is explained in detail in .. To avoid this problem, we use the homogeneous Faddeev equation Eq.(104)\[Fig.7\] to iterate the Faddeev amplitude in the final state. The diagram Fig.8.(a) is thus exactly equivalent to Fig.8.(a’), which is free from the additional delta function. We therefore have to evaluate the diagrams of Fig.8.(a’) and Fig.8.(b). Since the operator insertion on a constituent quark line has been given in Eqs. (42) – (47) and Fig.2.(a), we are led to the four diagrams shown in Fig.8.(d). Here, we used the identity Eq.(158) to express the matrix element only in terms of the Faddeev amplitudes in the rest frame, and used also the identity $$\gamma ^\mu (\mathrm{\Lambda }p)_\mu =S(\mathrm{\Lambda })\overline{)}pS(\mathrm{\Lambda }^1).$$ (73) Note that, since we truncated the $`qq`$ interaction to the scalar diquark channel, two of the boost matrices $`\widehat{S}(\mathrm{\Lambda }^1)`$ out of the three ($`\widehat{S}(\mathrm{\Lambda }^1)^3`$) in Eq.(158) cancel. The remaining $`\widehat{S}(\mathrm{\Lambda }^1)`$ is found in Fig.8.(d) at the point where the operator is inserted, which is a consequence of the manipulation Eq.(73). By using the following relations, which express the first order deviations due to the non-vanishing momentum transfer $`q`$: $`\delta \left(\text{}S(\mathrm{\Lambda }^1)\right)`$ $`=`$ $`S(\lambda )`$ (74) $`\delta \left(\text{}\gamma ^\mu (\mathrm{\Lambda }p)_\mu \right)`$ $`=`$ $`\left[\text{}S(\lambda ),\overline{)}p\right]`$ $`\delta \left(\text{}S_F(P\mathrm{\Lambda }^1(P/2l))\right)`$ $`=`$ $`S_F(P/2+l)\left[\text{}S(\lambda ),\overline{)}P/2\overline{)}l\right]S_F(P/2+l),\text{etc.},`$ we can consider the limit $`q0`$. After using the homogeneous Faddeev equation Eq.(104), we are left with the eight diagrams depicted in Fig.9. Note that the equality $`\mathrm{Fig}.\text{8}.(a)+\mathrm{Fig}.\text{8}.(b)(\text{}=\mathrm{Fig}.\text{8}.(\mathrm{d}))=\mathrm{Fig}.\text{9}`$ holds only in the limit $`q0`$, which is emphasized in Fig.9 by “$`q0`$”. By comparing both sides of Eq.(63) in the limit $`q0`$, the explicit spin-parity projection<sup>14</sup><sup>14</sup>14The spin-parity projection is performed by using the helicity formalism. (See for detail) shows that the first two diagrams are proportional to $`\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\gamma ^\mu \gamma _5u^{(\beta )b}(\stackrel{}{P})`$, which contribute to $`g_A`$, and the other diagrams are proportional to $`\overline{u}^{(\alpha )a}(\stackrel{}{P}){\displaystyle \frac{\tau ^b}{2}}\gamma _5S(\lambda )u^{(\beta )b}(\stackrel{}{P})`$, which contribute to $`g_{\pi NN}`$. Note that the 1st, the 3rd,the 4th and the 5th diagrams in Fig.9 come from the quark current contribution\[Fig.8.(a)\], and the others come from the exchange current contribution\[Fig.8.(b)\]. ### 5.2 Numerical Results in the Chiral Limit We first have to explain the choice of the parameters. There are four parameters: the cutoff $`\mathrm{\Lambda }`$ (Euclidean sharp cutoff), the current quark mass $`m_0`$, the effective coupling constant in the pionic channel $`g_\pi `$ and the effective coupling constant in the $`qq`$ scalar diquark channel $`g_{\mathrm{sd}}`$. We fix the first three parameters ($`\mathrm{\Lambda },m_0,g_\pi `$) by solving the gap equation Eq.(118) for the constituent quark mass $`M`$, and Eq.(124) for the pion mass $`m_\pi `$ and decay constant $`f_\pi `$, imposing the following three conditions: (1) $`m_\pi =140`$ MeV, (2) $`f_\pi =93`$ MeV, (3) $`M=400`$ MeV. The resulting values are $`\mathrm{\Lambda }=739`$ MeV, $`g_\pi =10.42`$ GeV<sup>-2</sup>, $`m_0=8.99`$ MeV. Once these parameters are fixed, we consider the chiral limit by taking $`m_00`$ keeping $`\mathrm{\Lambda }`$, $`g_\pi `$ fixed. Eq.(118) provides the $`m_0`$ dependences of $`M=M(m_0)`$. We treat $`g_{\mathrm{sd}}`$ as a free parameter (independent of $`m_0`$) to generate different nucleon masses. Note that, in the case $`m_\pi =140`$ MeV, $`g_{\mathrm{sd}}/g_\pi =0.66`$ gives the experimental value of the nucleon mass $`m_N=940`$ MeV. For convenience, we plot the nucleon mass $`m_N`$ (for $`g_{\mathrm{sd}}/g_\pi =0.66`$), the quark-diquark threshold($`m_N+m_{\mathrm{sd}}`$) where $`m_{\mathrm{sd}}`$ is the scalar diquark mass, the pion mass $`m_\pi `$, the pion decay constant $`f_\pi `$, and the constituent quark mass $`M`$ against the current quark mass $`m_0`$ in Fig.10. We also plot the nucleon mass in the chiral limit against $`g_{\mathrm{sd}}/g_\pi `$ \[solid line\], together with the physical case($`m_\pi =140`$ MeV) \[dashed line\] in Fig.11. Our results for $`g_{\pi NN}`$ and $`g_A`$ in the chiral limit are plotted against $`g_{\mathrm{sd}}/g_\pi `$ by the solid line in Fig.12 and Fig.13, respectively. We obtain $`g_{\pi NN}=13.2`$ and $`g_A=1.32`$ for $`g_{\mathrm{sd}}/g_\pi =0.66`$ compared to the experimental values $`g_{\pi NN}^{(\mathrm{exp})}=13.5`$ and $`g_A^{(\mathrm{exp})}=1.26`$. We will discuss the extension off the chiral limit in the next subsection. To estimate the violation of the GT relation, it is convenient to define a quantity: $$\mathrm{\Delta }_\mathrm{G}\frac{f_\pi g_{\pi NN}}{m_Ng_A}.$$ (75) In the chiral limit, the deviation of $`\mathrm{\Delta }_\mathrm{G}`$ from $`1`$ is solely due to the cutoff artifact. We plot $`\mathrm{\Delta }_\mathrm{G}`$ against $`g_{\mathrm{sd}}/g_\pi `$ in Fig.14 in the chiral limit \[solid line\]. It is seen that the violation of the GT relation is up to 4 %. In particular, for the reasonable case $`g_{\mathrm{sd}}/g_\pi =0.66`$, the violation is only 2 % in the chiral limit. ### 5.3 PCAC Violation off the Chiral Limit In the chiral limit, $`\mathrm{\Delta }_\mathrm{G}`$ works as the measure of the violation of the PCAC relation due to the cutoff artifact. However, off the chiral limit, $`\mathrm{\Delta }_\mathrm{G}`$ does not work as the measure any more, since it contains, in addition to the unphysical cutoff artifact which we are going to estimate here, the physical effect of the non-vanishing current quark mass. Therefore, in order to estimate the cutoff artifact, we need to construct a quantity which picks up only the cutoff artifact for any values of $`m_0>0`$. To this end, we introduce another form factor $`i_A(q^2)`$ through the following relation: $$N_\alpha ^a(\stackrel{}{P})\left|\text{}i\overline{\psi }(x)\gamma _5\frac{\tau ^b}{2}\psi (x)\right|N_\beta ^c(\stackrel{}{L})=e^{iqx}\overline{u}_\alpha ^a(\stackrel{}{P})\left(i\gamma _5\frac{\tau ^b}{2}\right)u_\beta ^c(\stackrel{}{L})\times i_A(q^2).$$ (76) Eq.(66) generalizes in the following way: $$2m_Ng_A(q^2)+q^2h_A(q^2)=2m_0i_A(q^2).$$ (77) In particular, in the limit $`q^20`$, since now $`m_\pi 0`$, we have $$2m_Ng_A=2m_0i_A(q^2=0).$$ (78) Now we define a quantity $`\mathrm{\Delta }_\mathrm{P}`$ as follows: $$\mathrm{\Delta }_\mathrm{P}=\frac{m_0i_A(q^2=0)}{m_Ng_A}.$$ (79) By construction, the deviation of $`\mathrm{\Delta }_\mathrm{P}`$ from 1 is solely due to the cutoff artifact for any value of $`m_0`$. To evaluate $`i_A(q^2)`$, we first have to solve the following equation for the operator insertion on a constituent quark line similar to Eq.(42): $`\left[{\displaystyle \frac{\delta S_F^{[e]}(x,y)}{\delta p^b(z)}}\right]^{[0]}`$ $`=`$ $`\begin{array}{c}S_F^{[0]}(x,z)\left(\text{}i\gamma _5{\displaystyle \frac{\tau ^b}{2}}\right)S_F^{[0]}(z,y)\hfill \\ +{\displaystyle d^4z^{}S_F^{[0]}(x,z^{})\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta p^b(z)}\right]^{[0]}S_F^{[0]}(z^{},y)}.\hfill \end{array}`$ (82) $`\left[{\displaystyle \frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta p^b(z)}}\right]^{[0]}`$ $`=`$ $`{\displaystyle \underset{\mathrm{\Gamma }}{}}2ig_\mathrm{\Gamma }^{(q\overline{q})}\text{Tr}\left(\left[{\displaystyle \frac{\delta S_F^{[e]}(z^{},z^{})}{\delta p^b(z)}}\right]^{[0]}\mathrm{\Gamma }\right)\mathrm{\Gamma }.`$ Since $`g_{\mathrm{ax}}=0`$, we can parameterize $`\delta \mathrm{\Sigma }/\delta p^b`$ as $$d^4z^{}e^{iq(z^{}z)}\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta p^b(z)}\right]=I(q^2)\left(\text{}i\gamma _5\frac{\tau ^b}{2}\right).$$ (83) $`I(q^2)`$ then satisfies the following equation: $$I(q^2)=2ig_\pi \mathrm{\Pi }_{55}(q^2)2ig_\pi \mathrm{\Pi }_{55}(q^2)I(q^2),$$ (84) with the solution $$I(q^2)=\frac{2ig_\pi \mathrm{\Pi }_{55}(q^2)}{1+2ig_\pi \mathrm{\Pi }_{55}(q^2)}.$$ (85) Note that this quantity of course has the pion pole. The evaluation of $`i_A(q^2=0)`$ amounts to the 2nd to the 8th diagrams given in Fig.9, where $`q_\mu \stackrel{~}{H}(q^2)`$ is replaced by $`1+I(q^2)`$. In Fig.15, we plot $`\mathrm{\Delta }_\mathrm{P}`$ against the current quark mass $`m_0`$ for $`g_{\mathrm{sd}}/g_\pi =0.66`$ case \[dashed line\]. It is seen that the cutoff artifact is again within 4%. ### 5.4 The Off-shell $`g_{\pi NN}`$ off the Chiral Limit We next evaluate $`g_{\pi NN}`$ off the chiral limit ($`m_\pi =140`$ MeV). In this case, whereas the definition of $`g_A=g_A(q^2=0)`$ does not change<sup>15</sup><sup>15</sup>15The evaluation of the iso-vector $`g_A`$ off the chiral limit was already done in ref. in the case of $`g_{\mathrm{ax}}=0`$., since the pion pole of $`h_A(q^2)`$ is shifted, $`g_{\pi NN}`$ is extracted according to: $$h_A(q^2)=\frac{2f_\pi g_{\pi NN}}{q^2m_\pi ^2}+\mathrm{}.$$ (86) In order to extract $`g_{\pi NN}`$ from the nucleon matrix element of the axial current, it is necessary to evaluate the form factor $`h_A(q^2)`$. However, at this stage it is still difficult to evaluate $`h_A(q^2)`$ for non-vanishing momentum transfer in the relativistic Faddeev approach<sup>16</sup><sup>16</sup>16The numerical evaluation of the on-shell $`g_{\pi NN}`$ off the chiral limit is currently under consideration.. We confine ourselves to evaluate the off-shell $`g_{\pi NN}(q^2=0)\stackrel{~}{g}_{\pi NN}`$ defined by the single-pole dominance approximation to $`h_A(q^2)`$ as $$\stackrel{~}{g}_{\pi NN}\frac{m_\pi ^2}{2f_\pi }h_A(q^2=0).$$ (87) Note that, whereas the nearest cut in the physical $`h_A(q^2)`$ is the three pion cut ($`q^2>9m_\pi ^2`$), the Cutokosky rule(cf. p.315 in ) suggests that the nearest cut in our $`h_A(q^2)`$ is the $`q\overline{q}`$ cut ($`q^2>4M^2`$). This is mainly due to the mean field approximation for the vacuum, and due to also the leak of the confinement in the NJL model. Since $`4M^2`$ is larger than $`9m_\pi ^2`$, the single-pole dominance approximation may work better in our case. We note that, although we could subtract the $`q\overline{q}`$ cut contributions from $`\stackrel{~}{H}(q^2)`$, it is impossible to subtract it from the remaining part, because the calculation refers only to $`q=0`$. The explicit evaluation shows that the subtraction of the $`q\overline{q}`$ cut contributions from $`\stackrel{~}{H}(q^2)`$ leads to only a small difference. We note that the quantity $`\stackrel{~}{g}_{\pi NN}`$ is not only one of the possible off-shell extensions of the on-shell $`g_{\pi NN}`$, but also the value of the axial form factor $`h_A(q^2)`$ at $`q^2=0`$ up to the well-defined numerical factor given by Eq.(87). This enables us to derive an analytical expression of $`\mathrm{\Delta }_\mathrm{G}`$ which follows from the PCAC relation by neglecting the small cutoff artifact. (See Appendix D.) We plot $`\stackrel{~}{g}_{\pi NN}`$ against $`g_{\mathrm{sd}}/g_\pi `$ in Fig.12 \[dashed line\], and $`g_A`$ in Fig.13 \[dashed line\]. The value of $`\stackrel{~}{g}_{\pi NN}`$ is 13.5 for $`g_{\mathrm{sd}}/g_\pi =0.66`$. This value is quite reasonable compared to the experimental value. We also plot $`\mathrm{\Delta }_\mathrm{G}`$ off the chiral limit \[dotted line\] and $`\mathrm{\Delta }_\mathrm{P}`$ \[dashed line\] against $`g_{\mathrm{sd}}/g_\pi `$ in Fig.14. The reader might suspect why the validity of the GT relation could be improved by going off the chiral limit. The reason is that the effect of non-vanishing $`m_0`$ works into the opposite direction compared to the UV-cutoff artifact. To see this, we plot $`\mathrm{\Delta }_\mathrm{G}`$ against the current quark mass $`m_0`$ \[solid line\] in Fig.15 for the case $`g_{\mathrm{sd}}/g_\pi =0.66`$. We also plot $`\mathrm{\Delta }_\mathrm{P}`$ against $`m_0`$ \[dashed line\], which is used to indicate the size of the cutoff artifact contained in $`\mathrm{\Delta }_\mathrm{G}`$. It is seen that $`\mathrm{\Delta }_\mathrm{G}`$ is a monotonically decreasing function of $`m_0`$. However, the experimental data $`\mathrm{\Delta }_\mathrm{G}=1.06`$ suggests that, as far as we believe that $`m_\pi =140`$ MeV is really close to the chiral limit, $`\mathrm{\Delta }_\mathrm{G}`$ should be an increasing function in the vicinity of $`m_00`$. We investigate this problem further in Appendix D by taking into account also the effects of non-vanishing coupling constant in the isovector axial vector $`q\overline{q}`$ channel $`g_{\mathrm{ax}}`$. (We leave this analysis for the appendix, because the non-vanishing $`g_{\mathrm{ax}}`$ makes things quite complicated.) The main conclusions there are summarized as follows: (1) $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ both decrease with increasing $`g_{\mathrm{ax}}/g_\pi `$, and increase with increasing $`m_0`$. (2) The best fit of $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ could be obtained in the region $`0g_{\mathrm{ax}}/g_\pi <0.1`$, which, however, would depend on the quantity which we prefer to adjust. From this point of view, $`g_{\mathrm{ax}}=0`$ is actually a rather good choice, because $`\stackrel{~}{g}_{\pi NN}`$ is very reasonable and $`g_A=1.33`$ is still close to the experimental value $`g_A^{(\mathrm{exp})}=1.26`$. (3) For those values of $`g_{\mathrm{ax}}`$ which we examine, $`\mathrm{\Delta }_\mathrm{G}`$ remains to be a decreasing function of $`m_0`$. (4) An analytic expression of $`\mathrm{\Delta }_\mathrm{G}`$ (Eq.(D.1)) is derived by neglecting the small cut-off artifact and by assuming that the vacuum is approximated by the mean-field approximation. All the baryonic quantities disappear from this expression. In particular, this expression is valid even beyond the ladder truncation scheme for the 3qBS kernel. The discrepancy between the analytic and the numeric $`\mathrm{\Delta }_\mathrm{G}`$ is due to the cutoff artifact. It is found to be within 3 %. (5) As a consequence, to make $`\mathrm{\Delta }_\mathrm{G}`$ to be an increasing function of $`m_0`$ and to obtain $`\mathrm{\Delta }_\mathrm{G}=1.06`$, we have to go beyond the validity of this analytic expression of $`\mathrm{\Delta }_\mathrm{G}`$. Therefore, all we can do is either to improve the gap equation for the vacuum beyond the mean field approximation or to estimate the on-shell $`g_{\pi NN}`$. We do not further investigate this problem in this paper. ## 6 Summary and Discussions In this work we reviewed how to evaluate expectation values of quantum one-body operators in the framework of the 3qBS /Faddeev equation by introducing classical external fields as a technical tool. In the 3qBS approach, the expectation values are obtained by sandwiching, the functional derivative of the 3qBS kernel with respect to the corresponding external field between the 3qBS amplitudes. In the Faddeev approach, the expectation values are obtained by sandwiching, between the Faddeev amplitudes, the functional derivative of the Faddeev kernel with respect to the corresponding external field. We gave the criterion for 3qBS kernels to give rise to the PCAC relation correctly. For practical purpose, we also gave the sufficient condition for the validity of this criterion by introducing the local “axial” gauge transformation of the external fields. We applied the sufficient condition to several 3qBS kernels. The main results are as follows: (1) If the 3qBS kernel is truncated in the ladder truncation scheme keeping only the $`qq`$ interaction in the scalar diquark channel, the PCAC relation is obtained correctly. (2) If the $`qq`$ interaction in the axial-vector diquark channel is included, it is necessary to include also the $`qq`$ interaction in the vector diquark channel to give rise to the PCAC relation correctly. (3) Even if the $`qq`$ interaction due to the $`q\overline{q}`$ exchange in both the pionic and the sigma mesonic channel is included, the correct PCAC relation is obtained. Concerning the point (2), we note that the vector diquark channel is often considered to be not important. This is because the non-relativistic quark model suggests that the contribution from this channel to the nucleon mass is suppressed in the non-relativistic limit. However, to respect the chiral $`SU(2)_L\times SU(2)_R`$ symmetry, the $`qq`$ interaction in the vector diquark channel should be included, even if it is expected to give a negligible contribution to the nucleon mass. We should note, however, that the relativistic Faddeev equation including all the $`qq`$ interactions in the scalar, axial-vector and vector diquark channel amounts to a two-dimensional integral equation with $`14\times 14`$ matrix structure even after the spin-parity projection is carried out, which requires a tremendous effort to be solved. Although these truncation schemes give rise to the PCAC relation correctly, the regularization scheme, which has been adopted in practical numerical calculations so far, does not respect the chiral symmetry, leading to the violation of the Goldberger-Treiman relation. To estimate this violation, we carried out the numerical evaluations of $`g_A`$ and $`g_{\pi NN}`$ in the chiral limit in the simplest case, i.e., the ladder truncation scheme keeping only the $`qq`$ interaction in the scalar diquark channel. We found that the violations are up to 4 %. In particular, for the case $`g_\mathrm{s}/g_\pi =0.66`$, which reproduces the experimental nucleon mass, the violation is only 2 %. The value of $`g_{\pi NN}`$ in the chiral limit is $`13.2`$ which is quite close to the experimental value $`13.5`$, and $`g_A`$ becomes $`1.32`$ compared to the experimental value $`1.26`$. We next estimated the PCAC violation due to the cutoff artifact off the chiral limit. We found that this violation is again within 4%. In the relativistic Faddeev method, it is still difficult to calculate form factors for non-vanishing momentum transfer, which is needed to extract the on-shell $`g_{\pi NN}`$ off the chiral limit. So we defined the off-shell $`\stackrel{~}{g}_{\pi NN}`$ by means of the single-pole dominance approximation to $`h_A(q^2)`$. Although we obtained a very reasonable result $`\stackrel{~}{g}_{\pi NN}=13.5`$ for the case $`g_{\mathrm{sd}}/g_\pi =0.66`$, the effect of non-vanishing current quark mass $`m_0`$ on the Goldberger-Treiman violation ($`\mathrm{\Delta }_\mathrm{G}(m_0)`$) was found to be in the “wrong” direction: Whereas the experimental value of $`\mathrm{\Delta }_\mathrm{G}`$ is $`1.06`$, which suggests that $`\mathrm{\Delta }_\mathrm{G}`$ should be an increasing function of $`m_0`$ in the vicinity of $`m_0=0`$, our $`\mathrm{\Delta }_\mathrm{G}(m_0)`$ is a decreasing function of $`m_0`$. We tried to resolve this problem (Appendix D) by taking into account the effect of non-vanishing effective coupling constant in the iso-vector axial-vector mesonic channel $`g_{\mathrm{ax}}`$. The main results are as follows: (1) $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ both decrease with increasing $`g_{\mathrm{ax}}/g_\pi `$, and increase with increasing $`m_0`$. (2) The best fit of $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ could be obtained in the region $`0g_{\mathrm{ax}}/g_\pi <0.1`$, which, however, would depend on the quantity which we prefer to adjust. From this point of view, $`g_{\mathrm{ax}}=0`$ is actually a rather good choice, since $`\stackrel{~}{g}_{\pi NN}=13.5`$ is a very reasonable result and $`g_A=1.33`$ is still close to the experimental value $`g_A^{(\mathrm{exp})}=1.26`$. (3) For those values of $`g_{\mathrm{ax}}`$ which we examined, $`\mathrm{\Delta }_\mathrm{G}`$ remains to be a decreasing function of $`m_0`$. (4) An analytic expression of $`\mathrm{\Delta }_\mathrm{G}`$ was derived by neglecting the small cutoff artifact and by assuming that the vacuum is approximated by the mean-field method. All the baryonic quantities disappear from this expression. In particular, this expression is valid even beyond the ladder truncation scheme for the 3qBS kernel. The discrepancy between the analytic and the numeric $`\mathrm{\Delta }_\mathrm{G}`$ is due to the cutoff artifact. It was found to be within 3 %. (5) As a consequence, to make $`\mathrm{\Delta }_\mathrm{G}`$ to be an increasing function of $`m_0`$, we have to take into account the effects which are beyond the validity of the analytic expression of $`\mathrm{\Delta }_\mathrm{G}`$. Hence, all we can do are either to improve the vacuum beyond the meanfield approximation or to estimate the on-shell $`g_{\pi NN}`$. We finally give a comment on the iso-scalar $`g_A^{(0)}`$. It is straightforward to extend our formalism to the chiral $`U(1)_L\times U(1)_R`$ case. The $`U_A(1)`$ anomaly in QCD is simulated in the NJL model as an explicit $`U_A(1)`$ symmetry breaking. It is easy to see that, whereas the vector, axial-vector and tensor diquark channels form closed chiral $`U(1)_L\times U(1)_R`$ singlets separately, only a combination of the scalar diquark and the pseudo scalar diquark forms a closed chiral $`U(1)_L\times U(1)_R`$ doublet. To isolate the $`U_A(1)`$ breaking contribution, one can parameterize the two coupling constants $`g_{\mathrm{sd}}`$ and $`g_{\mathrm{pd}}`$ as $`g_{\mathrm{sd}}=\lambda +\delta \lambda `$, $`g_{\mathrm{pd}}=\lambda +\delta \lambda `$. Now, in the ladder truncation scheme, it is only $`\delta \lambda `$ that can provide the $`U_A(1)`$ breaking effects to 3qBS amplitudes, because $`\lambda `$ and the other $`qq`$ interactions respect the $`U_A(1)`$ symmetry. The $`qq`$ interaction in the pseudo-scalar diquark channel is often considered to be irrelevant from the non-relativistic analogy. However, setting $`g_{\mathrm{pd}}=0`$ corresponds to a particular choice of $`U_A(1)`$ breaking, i.e., $`\delta \lambda =g_{\mathrm{sd}}/2`$, and this particular choice is not based on any of the underlying physics of $`U_A(1)`$ breaking. Due to this reason, we suggest that the $`qq`$ interaction in the pseudoscalar diquark channel should be included for a reasonable estimate of the iso-scalar $`g_A^{(0)}`$, even if it is expected to give a negligible contribution to the nucleon mass. Acknowledgment The author thanks K. Yazaki, W. Bentz, H. Asami, L. v. Smekal and H. Terazawa for their extensive discussions and encouraging suggestions. He also thanks the unknown referee of his previous paper for giving him the main motivation for the current work. ## Appendices ## Appendix A The Faddeev Equation The aim of this appendix is to summarize the notations of the relativistic Faddeev equation and to provide the derivations of some of the relevant relations involving the Faddeev method which are essential to the other parts of this paper in a self-contained manner. In this appendix, summations over repeated indices are not implied, and, unless explicitly indicated, the external fields are understood to be absent, i.e., $`va0`$, $`mm_0`$. ### A.1 The Relativistic Faddeev Equation<sup>17</sup><sup>17</sup>17A rather good pedagogical introduction to the Faddeev equation is found in . and the Green’s Function We begin with the 3qBS equation (see Eq.(4)) in the absence of the external fields: $$G=G_0+KG;K=K_1+K_2+K_3,$$ (88) where the index $`i`$ of $`K_i`$ refers to the spectator quark. The formal solution of $`G`$ is expressed by using the resolvent of $`K`$ as follows: $$G=\frac{1}{1K}G_0.$$ (89) Note that the resolvent exists in a mathematical sense. However, because $`K`$ is an unbounded operator, it is difficult to interpret it as it stands. Therefore, we adopt the Faddeev prescription. We introduce the following Faddeev decomposition of the Green’s function: $$G=G_0+G^1+G^2+G^3;G^iK_iG.$$ (90) We insert the following resolvent identity of $`K`$ into Eq.(89): $$\frac{1}{1K}=\frac{1}{1K_i}+\frac{1}{1K_i}\left(\text{}\underset{ji}{}K_j\right)\frac{1}{1K}.$$ (91) We obtain $`G={\displaystyle \frac{1}{1K_i}}G_0+{\displaystyle \frac{1}{1K_i}}{\displaystyle \underset{ji}{}}G^j,`$ (92) which is further inserted into the defining relation of $`G^i`$ in Eq.(90). We are left with the following closed equations for $`G^1,G^2,G^3`$ (the Faddeev equations): $$G^i=\frac{K_i}{1K_i}G_0+\frac{K_i}{1K_i}\underset{ji}{}G^j.$$ (93) It is possible to simplify these coupled equations into a single closed equation for $`G^3`$ by using the identical particle nature of the three quarks. $`G^1`$ and $`G^2`$ are obtained from $`G^3`$ by means of simple permutation operations. For this purpose, it is convenient to introduce the cyclic permutation operator $`Z`$, which is defined as follows: $$(Z\psi )(x_1,x_2,x_3)=\psi (x_2,x_3,x_1).$$ (94) Note that $`Z`$ is easily implemented by using delta functions. We give a list of obvious relations: $$\begin{array}{c}Z^3=1,Z(1+Z+Z^2)=(1+Z+Z^2)Z=1+Z+Z^2,\hfill \\ \text{}K_i=Z^iK_3Z^i,ZG=GZ=G,ZG_0=G_0Z=G_0,G^i=Z^iG^3.\hfill \end{array}$$ (95) Now Eq.(93) reduces to the following closed integral equation for $`G^3`$ (the reduced Faddeev equation): $$G^3=\frac{K_3}{1K_3}G_0+\frac{K_3}{1K_3}(Z+Z^2)G^3,$$ (96) which is depicted in Fig.7.(a). The two-quark resolvent $`{\displaystyle \frac{K_3}{1K_3}}`$ is depicted in Fig.7.(b). (cf. Eq.(5)) It is not so hard to identify the so-called “Z-diagram” structure (the quark exchange diagram), which is provided by a combination of the permutation operator $`Z`$ (or $`Z^2`$) and two external quark propagators of the two-quark resolvent $`{\displaystyle \frac{K_3}{1K_3}}`$. We refer to the kernel of this integral equation as the Faddeev kernel $`K_F`$: $$K_F\frac{K_3}{1K_3}(Z+Z^2).$$ (97) Now the formal solution of $`G^3`$ to Eq.(96) is given as: $$G^3=\frac{1}{1K_F}\frac{K_3}{1K_3}G_0.$$ (98) By inserting this into the first relation in Eq.(90), we have another formal representation of the Green’s function $`G`$: $$G=G_0+(1+Z+Z^2)\frac{1}{1K_F}\frac{K_3}{1K_3}G_0.$$ (99) ### A.2 The 3qBS Amplitude and the Faddeev Amplitude To obtain the form of the Green’s function $`G`$ near the nucleon pole, we first diagonalize the Faddeev kernel $`K_F`$, regarding the total four momentum $`p_\mu `$ as a parameter<sup>19</sup><sup>19</sup>19The numerical procedure to solve this homogeneous Faddeev equation is explained into detail in .: $$K_F\psi [n;p]=\lambda _n(p^2)\psi [n;p],\overline{\psi }[n;p]K_F=\lambda _n(p^2)\overline{\psi }[n;p].$$ (100) We adopt the following normalization condition: $$\overline{\psi }[n^{};p^{}]\psi [n;p]=𝒩_n(p^2)(2\pi )^4\delta ^{(4)}(p^{}p)\delta _{n^{}n};𝒩_n(p^2)\frac{i}{2\sqrt{p^2}\lambda _n^{}(p^2)},$$ (101) where $`\lambda _n^{}(p^2)d\lambda _n(p^2)/d(p^2)`$. Now the Faddeev kernel is expressed as: $$K_F=\underset{n}{}\frac{d^4p}{(2\pi )^4𝒩_n(p^2)}\lambda _n(p^2)\psi [n;p]\overline{\psi }[n;p].$$ (102) Note that, in reasonable cases, the eigen-modes associated with the largest eigenvalue are four-fold degenerate —they correspond to the nucleon (the ground states in the sector of the baryon number 1). The degeneracy is due to the iso-spin 1/2 and spin up/down. We refer to these eigen-modes as the (off-shell) eigenvalues and Faddeev amplitudes for the nucleon and denote them as $`\lambda _N(p^2)`$, $`\psi [N_\alpha ^a(p)]`$ and $`\overline{\psi }[N_\alpha ^a(p)]`$. The nucleon mass $`m_N`$ is obtained by solving the following equation: $$\lambda _N(p^2=m_N^2)=1.$$ (103) The associated eigenvectors satisfy the homogeneous Faddeev equations: $$\psi [N_\beta ^c(\stackrel{}{L})]=\frac{K_3}{1K_3}(Z+Z^2)\psi [N_\beta ^c(\stackrel{}{L})],\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]=\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]\frac{K_3}{1K_3}(Z+Z^2).$$ (104) For simplicity, we suppress the time components of the total four momenta in the on-shell Faddeev amplitudes. We define the 3qBS amplitudes for nucleon states by: $`\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]`$ $``$ $`(1+Z+Z^2)\psi [N_\alpha ^a(\stackrel{}{P})]`$ (105) $`\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{P})]`$ $``$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{K_3}{1K_3}}G_0`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{K_3}{1K_3}}{\displaystyle \frac{1+Z+Z^2}{3}}G_0.`$ In order to see that these definitions of the 3qBS amplitudes are reasonable and that the normalization scheme adopted in Eq.(101) is consistent with the covariant normalization of the ket vectors in Eq.(11), we first consider the form of the Green’s function near the nucleon pole as follows (cf. Eq.(99)): $`G`$ $``$ $`(1+Z+Z^2){\displaystyle \underset{a,\alpha }{}}{\displaystyle \frac{d^4P}{(2\pi )^4𝒩_N(P^2)}\frac{\psi [N_\alpha ^a(P)]\overline{\psi }[N_\alpha ^a(P)]}{1\lambda _N(P^2)}\frac{K_3}{1K_3}G_0}`$ $``$ $`{\displaystyle \underset{a,\alpha }{}}{\displaystyle \frac{d^4P}{(2\pi )^4𝒩_N(P^2)}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]\frac{1}{1\lambda _N(P^2)}\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{P})]}`$ $``$ $`{\displaystyle \underset{a,\alpha }{}}{\displaystyle \frac{d^4P}{(2\pi )^4𝒩_N(m_N^2)}\times \frac{1}{\lambda _N^{}(m_N^2)}\frac{\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{P})]}{P^2m^2+iϵ}}`$ $``$ $`{\displaystyle \underset{a,\alpha }{}}{\displaystyle \frac{d^4P}{(2\pi )^4}\frac{m_N}{E_N(\stackrel{}{P}^2)}\frac{i}{P_0E_N(\stackrel{}{P}^2)+iϵ}\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]\overline{\mathrm{\Psi }}[N_\alpha ^a(\stackrel{}{P})]}.`$ This result is consistent with the expression which is expected from the canonical operator analysis. Note that the above manipulations are exact at the nucleon pole. Next, we show that these 3qBS amplitudes satisfy the homogeneous 3qBS equation Eq.(17): $`K\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]`$ $`=`$ $`{\displaystyle \underset{i=1,2,3}{}}Z^iK_3Z^i(1+Z+Z^2)\psi [N_\alpha ^a(\stackrel{}{P})]`$ $`=`$ $`(1+Z+Z^2)K_3(1+Z+Z^2)\psi [N_\alpha ^a(\stackrel{}{P})]`$ $`=`$ $`(1+Z+Z^2){\displaystyle \frac{K_3}{1K_3}}(Z+Z^2)\psi [N_\alpha ^a(\stackrel{}{P})]=\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})]`$ $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{P})]K`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{K_3}{1K_3}}{\displaystyle \frac{1+Z+Z^2}{3}}{\displaystyle \underset{i=1,2,3}{}}Z^iK_3Z^i`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{K_3}{1K_3}}(1+Z+Z^2)K_3{\displaystyle \frac{1+Z+Z^2}{3}}`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{K_3}{1K_3}}{\displaystyle \frac{1+Z+Z^2}{3}}=\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{P})],`$ where, to obtain the 3rd lines of Eq.(A.2) and Eq.(A.2), we used the following identities: $$\begin{array}{c}\overline{\psi }\frac{K_3}{1K_3}(1+Z+Z^2)=\overline{\psi }\frac{K_3}{1K_3}+\overline{\psi }=\overline{\psi }\frac{1}{1K_3}\hfill \\ \text{}(1+Z+Z^2)\psi =\frac{K_3}{1K_3}(Z+Z^2)\psi +(Z+Z^2)\psi =\frac{1}{1K_3}(Z+Z^2)\psi ,\hfill \end{array}$$ (109) which immediately follow from the homogeneous Faddeev equations Eq.(104). ### A.3 The Matrix Elements in terms of the Faddeev Amplitude Here, we derive an explicit expression for the matrix element of the axial vector current operator in terms of the Faddeev amplitudes, which is, in the practical applications, more convenient than the expression in terms of the 3qBS amplitudes. To avoid cumbersome notations, we introduce a shorthand notation: $`O_{i;\mu }^b\left[{\displaystyle \frac{i\delta K_i^{[e]}}{\delta a_\mu ^b(x=0)}}\right]^{[0]}`$ (i=1,2,3) and $`O_\mu ^b={\displaystyle \underset{i=1,2,3}{}}O_{i;\mu }^b`$, where the index $`i`$ refers to the spectator quark. Note that $`O_{i;\mu }^b=Z^iO_{3;\mu }^bZ^i`$. Now we have from Eq.(21): $`N_\alpha ^a(\stackrel{}{P})\left|\text{}A_\mu ^b(x=0)\right|N_\beta ^c(\stackrel{}{L})`$ (110) $`=`$ $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(\stackrel{}{P})]O_\mu ^b\mathrm{\Psi }[N_\beta ^c(\stackrel{}{L})]`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{K_3}{1K_3}}(1+Z+Z^2)O_{3;\mu }^b(1+Z+Z^2)\psi [N_\beta ^c(\stackrel{}{L})]`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]{\displaystyle \frac{1}{1K_3}}O_{3;\mu }^b{\displaystyle \frac{1}{1K_3}}(Z+Z^2)\psi [N_\beta ^c(\stackrel{}{L})]`$ $`=`$ $`\overline{\psi }[N_\alpha ^a(\stackrel{}{P})]\left[{\displaystyle \frac{i\delta K_F^{[e]}}{\delta a_\mu ^a(x=0)}}\right]^{[0]}\psi [N_\beta ^c(\stackrel{}{L})],`$ where, to obtain the fourth line, we used Eq.(109). Since the permutation operator $`Z`$ does not depend on the external fields, $`\delta /\delta a_\mu ^b`$ only hits one the constituent quark propagators in the two-quark resolvent $`{\displaystyle \frac{K_3}{1K_3}}`$ in $`K_F`$. (cf. Eq.(97)) We comment on the ladder truncation scheme. In this case, the $`qq`$ interaction is point-like. By combining the diagramatic expressions of $`K_3`$ \[Fig.3.(a)\] and the two-quark resolvent $`{\displaystyle \frac{K_3}{1K_3}}`$ \[Fig.7.(b)\], we see that the diagrams involved in $`i\delta K_F/\delta a_\mu ^a`$ are classified into two types, i.e., (1) $`\delta /\delta a_\mu ^a`$ hits one of the internal quark propagators (i.e., in the $`qq`$ bubble diagram) in the ladder sum, (2) $`\delta /\delta a_\mu ^a`$ hits one of the two external quark propagators. With the aid of this classification, a straight forward diagramatic argument shows that the diagramatic expression of Eq.(110) is given by the three diagrams Fig.8.(a), Fig.8.(b) and Fig.8.(c). The first type leads to Fig.8.(c), which we refer to as the “diquark current” contribution, and the second type leads to Fig.8.(a) and Fig.8.(b), which we refer to as the “quark current” contribution and the “exchange current” contribution, respectively. ## Appendix B Notations of the NJL Model The aims of this section are to define some of the relevant quantities with explicit examples, and to summarize the notations of the NJL model. In this section, the external fields are understood to be absent, i.e., $`v_\mu ^a(x)a_\mu ^a(x)0`$, $`m(x)m_0`$. ### B.1 The Elementary Effective Coupling Constants and the Closed Chiral Multiplets Here we define the elementary effective coupling constants in $`q\overline{q}`$ and $`qq`$ channels. We start with the Lagrangian density Eq.(1). For definiteness, we virtually distinguish the two $`\psi `$’s and the two $`\overline{\psi }`$’s from each other, respectively, i.e., $$_I=\underset{\mathrm{\Gamma }}{}g_\mathrm{\Gamma }(\overline{\psi }_1\mathrm{\Gamma }\psi _2)(\overline{\psi }_3\mathrm{\Gamma }\psi _4).$$ (111) The elementary interaction, which is depicted in Fig.16.(a), can be classified into the following three types: (i) $`q\overline{q}`$ direct channel \[Fig.16.(b)\], (ii) $`q\overline{q}`$ exchange channel \[Fig.16.(c)\], (iii) $`qq`$ diquark channel \[Fig.16.(d)\]. They are related by Fierz identities to each other. We define the effective coupling constants $`g^{(q\overline{q}\mathrm{dir})}`$, $`g^{(q\overline{q}\mathrm{exch})}`$ and $`g^{(qq)}`$ through the following relations: $`_I`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \underset{i=0}{\overset{3}{}}}{\displaystyle \underset{A=0}{\overset{8}{}}}g_{\alpha iA}^{(q\overline{q}\mathrm{dir})}\left(\text{}\overline{\psi }_1\mathrm{\Gamma }_\alpha \tau _i\beta _A\psi _2\right)\left(\text{}\overline{\psi }_3\mathrm{\Gamma }^\alpha \tau _i\beta _A\psi _4\right)`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \underset{i=0}{\overset{3}{}}}{\displaystyle \underset{A=0}{\overset{8}{}}}g_{\alpha iA}^{(q\overline{q}\mathrm{exch})}\left(\text{}\overline{\psi }_1\mathrm{\Gamma }_\alpha \tau _i\beta _A\psi _4\right)\left(\text{}\overline{\psi }_3\mathrm{\Gamma }^\alpha \tau _i\beta _A\psi _2\right)`$ $`=`$ $`{\displaystyle \underset{\alpha }{}}{\displaystyle \underset{i=0}{\overset{3}{}}}{\displaystyle \underset{A=0}{\overset{8}{}}}g_{\alpha iA}^{(qq)}\left(\text{}\overline{\psi }_1(\mathrm{\Gamma }_\alpha \gamma _5C^1)(\tau _i\tau _2)\beta _A\overline{\psi }_3^T\right)\left(\text{}\psi _2^T(C\gamma _5\mathrm{\Gamma }^\alpha )(\tau _2\tau _i)\beta _A\psi _4\right).`$ $`\mathrm{\Gamma }_\alpha ,\mathrm{\Gamma }^\alpha `$ are Dirac gamma matrices, where $`\alpha `$ runs over S(scalar), V(vector), T(tensor), A(axial-vector) and P(pseudo-scalar), i.e., $`(\mathrm{\Gamma }_\alpha )_{\alpha =\mathrm{S},\mathrm{V},\mathrm{T},\mathrm{A},\mathrm{P}}`$ $`=`$ $`(1,\gamma _\mu ,\sigma _{\mu \nu },\gamma _\mu \gamma _5,\gamma _5)`$ (113) $`(\mathrm{\Gamma }^\alpha )_{\alpha =\mathrm{S},\mathrm{V},\mathrm{T},\mathrm{A},\mathrm{P}}`$ $`=`$ $`(1,\gamma ^\mu ,\sigma ^{\mu \nu },\gamma _5\gamma ^\mu ,\gamma _5).`$ $`\tau _i`$ ($`i=0,1,2,3`$) are the iso-spin Pauli matrices with $`\tau _0=1`$. They are normalized according to $`\text{tr}(\tau _i\tau _j)=2\delta _{ij}`$. Note that, both in $`q\overline{q}`$ and $`qq`$ representations, $`\tau _0`$ corresponds to the iso-scalar, and $`\tau _i`$ ($`i=1,2,3`$) correspond to the iso-vector channels. $`\beta _A`$ are the rescaled Gell-Mann color matrices, i.e., $`\beta _0=1`$, $`\beta _A=\sqrt{{\displaystyle \frac{3}{2}}}\lambda _A`$ ($`A=1,2,\mathrm{},8`$) with the normalization $`\text{tr}(\beta _A\beta _B)=3\delta _{AB}`$. In $`q\overline{q}`$ representations, $`\beta _0`$ corresponds to the $`1_c`$ mesonic channels, and $`\beta _A`$ for $`A=1,2,\mathrm{},8`$ corresponds to $`8_c`$ mesonic channels. Note that $`\beta _A`$ for $`A=2,5,7`$ are anti-symmetric and the others are symmetric. Therefore, in $`qq`$ representation, $`\beta _A`$ for $`A=2,5,7`$ correspond to $`\overline{3}_c`$ diquark channels, and $`\beta _A`$ for $`A=0,1,3,4,6,8`$ correspond to $`6_c`$ diquark channels. We define the effective coupling constants $`g^{(q\overline{q})}`$ as follows: $$g_{\alpha iA}^{(q\overline{q})}=g_{\alpha iA}^{(q\overline{q}\mathrm{dir})}+g_{\alpha iA}^{(q\overline{q}\mathrm{exch})}.$$ (114) We consider two examples (i) the original NJL type interaction Lagrangian: $$_I=g\left(\text{}(\overline{\psi }\psi )^2\underset{i=1}{\overset{3}{}}(\overline{\psi }\gamma _5\tau _i\psi )^2\right),$$ (115) where $`\tau _i`$ is the isospin Pauli matrix, and (ii) the color current type interaction Lagrangian: $$_I=g\underset{a=1}{\overset{8}{}}\left(\text{}\overline{\psi }\gamma _\mu \frac{\lambda _a}{2}\psi \right)^2,$$ (116) where $`\lambda _a`$ is the color Gell-Mann matrix. We give a list of these effective coupling constants $`g^{(q\overline{q})}=g^{(q\overline{q}\mathrm{dir})}+g^{(q\overline{q}\mathrm{exch})}`$ and $`g^{(qq)}`$ for these two interaction Lagrangians: | Original NJL Type | S | V | T | A | P | | --- | --- | --- | --- | --- | --- | | $`1_c`$ ($`q\overline{q}`$), $`I=0`$ | $`g+g/12`$ | $`0g/6`$ | $`0+g/12`$ | $`0g/6`$ | $`0+g/12`$ | | $`1_c`$ ($`q\overline{q}`$), $`I=1`$ | $`0g/12`$ | $`0+0`$ | $`0g/12`$ | $`0+0`$ | $`gg/12`$ | | $`8_c`$ ($`q\overline{q}`$), $`I=0`$ | $`0+g/12`$ | $`0g/6`$ | $`0+g/12`$ | $`0g/6`$ | $`0+g/12`$ | | $`8_c`$ ($`q\overline{q}`$), $`I=1`$ | $`0g/12`$ | $`0+0`$ | $`0g/12`$ | $`0+0`$ | $`0g/12`$ | | $`\overline{3}_c`$ ($`qq`$), $`I=0`$ | $`g/6`$ | $`g/12`$ | $`g/6`$ (!) | $`g/12`$ (!) | $`g/6`$ | | $`\overline{3}_c`$ ($`qq`$), $`I=1`$ | 0 (!) | $`g/12`$ (!) | 0 | $`g/12`$ | 0 (!) | | $`6_c`$ ($`qq`$), $`I=0`$ | $`g/6`$ (!) | $`g/12`$ (!) | $`g/6`$ | $`g/12`$ | $`g/6`$ (!) | | $`6_c`$ ($`qq`$), $`I=1`$ | 0 | $`g/12`$ | 0 (!) | $`g/12`$ (!) | 0 | | Color Current Type | S | V | T | A | P | | $`1_c`$ ($`q\overline{q}`$), $`I=0`$ | $`0+2g/9`$ | $`0g/9`$ | $`0+0`$ | $`0+g/9`$ | $`02g/9`$ | | $`1_c`$ ($`q\overline{q}`$), $`I=1`$ | $`0+2g/9`$ | $`0g/9`$ | $`0+0`$ | $`0+g/9`$ | $`02g/9`$ | | $`8_c`$ ($`q\overline{q}`$), $`I=0`$ | $`0g/36`$ | $`g+g/72`$ | $`0+0`$ | $`0g/72`$ | $`0+g/36`$ | | $`8_c`$ ($`q\overline{q}`$), $`I=1`$ | $`0g/36`$ | $`0+g/72`$ | $`0+0`$ | $`0g/72`$ | $`0+g/36`$ | | $`\overline{3}_c`$ ($`qq`$), $`I=0`$ | $`g/9`$ | $`g/18`$ | 0 (!) | $`g/18`$ (!) | $`g/9`$ | | $`\overline{3}_c`$ ($`qq`$), $`I=1`$ | $`g/9`$ (!) | $`g/18`$ (!) | 0 | $`g/18`$ | $`g/9`$ (!) | | $`6_c`$ ($`qq`$), $`I=0`$ | $`g/18`$ (!) | $`g/36`$ (!) | 0 | $`g/36`$ | $`g/18`$ (!) | | $`6_c`$ ($`qq`$), $`I=1`$ | $`g/18`$ | $`g/36`$ | 0 (!) | $`g/36`$ (!) | $`g/18`$ | For completeness, we also presented the effective couplings in the $`6_c`$ diquark channels, which do not contribute directly to the color singlet baryon states. Note that the two $`\overline{\psi }`$ fields and the two $`\psi `$ fields are originally undistinguished Grassmann fields. Therefore, in $`qq`$ representation, unless $`\mathrm{\Gamma }_\alpha \gamma _5C^1\tau _i\tau _2\beta _A`$ is anti-symmetric, the contribution vanishes. Due to this reason, half of the effective coupling constants in the $`qq`$ channels actually vanish, and these cases are indicated by “(!)” in the list. Due to the chiral $`SU(2)_L\times SU(2)_R`$ symmetry of the interaction Lagrangian, it can be already seen from the above list that some of the effective coupling constants are grouped together. In fact, the straight forward application of the chiral $`SU(2)_L\times SU(2)_R`$ transformation leads us to the following list of “closed chiral multiplets”<sup>20</sup><sup>20</sup>20For the precise meaning, see Section.4.: $$\begin{array}{cccc}& & & \\ & & & \\ \text{}q\overline{q}(1_c)\hfill & \left(\text{}V,I=0\right)\hfill & q\overline{q}(8_c)\hfill & \left(\text{}V,I=0\right)\hfill \\ \text{}\hfill & \left(\text{}A,I=0\right)\hfill & & \left(\text{}A,I=0\right)\hfill \\ \text{}\hfill & \left(\text{}(S,I=0)(P,I=1)\right)\hfill & & \left(\text{}(S,I=0)(P,I=1)\right)\hfill \\ \text{}\hfill & \left(\text{}(S,I=1)(P,I=0)\right)\hfill & & \left(\text{}(S,I=1)(P,I=0)\right)\hfill \\ \text{}\hfill & \left(\text{}(T,I=0)(T,I=1)\right)\hfill & & \left(\text{}(T,I=0)(T,I=1)\right)\hfill \\ \text{}\hfill & \left(\text{}(V,I=1)(A,I=1)\right)\hfill & & \left(\text{}(V,I=1)(A,I=1)\right)\hfill \\ & & & \\ & & & \\ \text{}qq(\overline{3}_c)\hfill & \left(\text{}S,I=0\right)\hfill & qq(6_c)\hfill & \left(\text{}S,I=1\right)\hfill \\ \text{}\hfill & \left(\text{}T,I=1\right)\hfill & & \left(\text{}T,I=0\right)\hfill \\ \text{}\hfill & \left(\text{}P,I=0\right)\hfill & & \left(\text{}P,I=1\right)\hfill \\ \text{}\hfill & \left(\text{}(V,I=0)(A,I=1)\right)\hfill & & \left(\text{}(V,I=1)(A,I=0)\right)\hfill \\ & & & \end{array}$$ Note that the total number of independent coupling constants of the chiral $`SU(2)_L\times SU(2)_R`$ symmetric NJL type interaction Lagrangian is at most eight. Therefore, if we fix all the eight coupling constants in the $`qq`$ channels, all the effective coupling constants in the $`q\overline{q}`$ channels are obtained as their linear combinations. In the ladder truncation scheme of the 3qBS kernel, we can fix the coupling constants in the $`\overline{3}_c`$ ($`qq`$) channels based on the calculations for color singlet baryons. However, the coupling constants in the $`6_c`$ ($`qq`$) channels remain free. One can then use these remaining coupling constants (related to the $`q\overline{q}`$ coupling constants) to reproduce the mesonic properties. To avoid cumbersome notations, we introduce the following abbreviations: $$g_\pi =g_{P,I=1,1_c}^{(q\overline{q})}=g_{S,I=0,1_c}^{(q\overline{q})},g_{\mathrm{sd}}=g_{S,I=0,\overline{3}_c}^{(qq)},g_{\mathrm{ax}}=g_{A,I=1,1_c}^{(q\overline{q})}=g_{V,I=1,1_c}^{(q\overline{q})}.$$ (117) ### B.2 The Vacuum of the NJL model at the Mean Field Level Here, we summarize some of the notations of the NJL model in the vacuum and the mesonic sectors. #### B.2.1 $`g_{\mathrm{ax}}=0`$ Case The gap equation is given by: $$M=m_0+2ig_\pi \frac{d^4p}{(2\pi )^4}\text{Tr}S_F(p),S_F(p)\frac{1}{\overline{)}pM},$$ (118) which provides the spontaneous breaking of the chiral symmetry in the NJL model at the level of mean field approximation. $`M`$ is the constituent quark mass. The pion mass and the pion decay constant are obtained from the two-point axial current correlator as follows: $`{\displaystyle d^4xe^{iqx}0\left|\text{}TA_\mu ^a(x)A_\nu ^b(0)\right|0}`$ $`=`$ $`{\displaystyle \frac{iq_\mu q_\nu f_\pi ^2}{q^2m_\pi ^2+iϵ}}+\text{continuum}.`$ (119) Here we adopted the following definition of the pion decay constant: $$0\left|\text{}A_\mu ^a(x)\right|\pi ^b(\stackrel{}{p})=ip_\mu f_\pi \delta _{ab}e^{ipx},$$ (120) where the covariant normalization $`\pi ^a(\stackrel{}{p})|\text{}\pi ^b(\stackrel{}{k})=2\sqrt{m_\pi ^2+\stackrel{}{p}^2}(2\pi )^3\delta ^{(3)}(\stackrel{}{p}\stackrel{}{k})\delta _{ab}`$ is adopted. The explicit form of the axial current correlator can be easily obtained in the ladder approximation. However, in order to emphasize the consistency between the vacuum, the mesonic sector and the baryonic sector in our formulation, we prefer to use the external field method. Since the canonical operator expression of $`\delta S_F/\delta a_\nu ^b`$ is $$\left[\frac{i\delta iS_F^{[e]}(x,z)_{\alpha \beta }}{\delta a_\nu ^b(y)}\right]^{[0]}=0\left|\text{}T\psi _\alpha (x)\overline{\psi }_\beta (z)A_\nu ^b(y)\right|0,$$ (121) the explicit form of the axial current correlator is obtained in the following way(cf. eqs. (44) - (47)): $`{\displaystyle d^4xe^{iq(xy)}0\left|\text{}TA_\mu ^a(x)A_\nu ^b(y)\right|0}`$ $`=`$ $`{\displaystyle d^4xe^{iq(xy)}\text{Tr}\left((\gamma _\mu \gamma _5\frac{\tau ^a}{2})\left[\frac{\delta S_F^{[e]}(x,x)}{\delta a_\mu ^b(y)}\right]^{[0]}\right)}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\delta _{ab}\mathrm{\Pi }_{\mu \nu }(q){\displaystyle \frac{1}{4}}q_\mu q_\nu \delta _{ab}\mathrm{\Pi }_{5A}(q^2)\stackrel{~}{H}(q^2)`$ $`=`$ $`{\displaystyle \frac{1}{4}}\mathrm{\Pi }_{\mu \nu }(q)\delta _{ab}{\displaystyle \frac{1}{4}}q_\mu q_\nu \delta _{ab}{\displaystyle \frac{\mathrm{\Pi }_{5A}(q^2)(2ig_\pi )\mathrm{\Pi }_{5A}(q^2)}{1+2ig_\pi \mathrm{\Pi }_{55}(q^2)}},`$ where $`\mathrm{\Pi }_{55}(q^2)\delta _{ab}`$ $``$ $`{\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}\left(\text{}(\gamma _5\tau ^a)iS_F(k+q)(\gamma _5\tau ^b)iS_F(k)\right)}`$ (123) $`q_\mu \mathrm{\Pi }_{5A}(q^2)\delta _{ab}`$ $``$ $`{\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}\left(\text{}(\gamma _5\tau ^a)iS_F(k+q)(\gamma _\mu \gamma _5\tau ^b)iS_F(k)\right)}`$ $`=`$ $`{\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}\left(\text{}(\gamma _5\gamma _\mu \tau ^b)iS_F(k+q)(\gamma _5\tau ^a)iS_F(k)\right)}`$ $`\mathrm{\Pi }_{\mu \nu }(q)\delta _{ab}`$ $``$ $`{\displaystyle \frac{d^4k}{(2\pi )^4}\text{Tr}\left(\text{}(\gamma _\mu \gamma _5\tau ^a)iS_F(k+q)(\gamma _5\gamma _\nu \tau ^b)iS_F(k)\right)}`$ $``$ $`\delta _{ab}\left({\displaystyle \frac{q_\mu q_\nu }{q^2}}\right)\mathrm{\Pi }_1(q^2)+\delta _{ab}\left(g_{\mu \nu }{\displaystyle \frac{q_\mu q_\nu }{q^2}}\right)\mathrm{\Pi }_2(q^2).`$ The explicit expressions for $`m_\pi `$ and $`f_\pi `$ are obtained from: $$0=1+2ig_\pi \mathrm{\Pi }_{55}(m_\pi ^2),f_\pi =\frac{i\mathrm{\Pi }_{5A}(m_\pi ^2)}{2\sqrt{i\mathrm{\Pi }_{55}^{}(m_\pi ^2)}},$$ (124) where $`\mathrm{\Pi }_{55}^{}(q^2){\displaystyle \frac{d\mathrm{\Pi }_{55}(q^2)}{d(q^2)}}`$. #### B.2.2 $`g_{\mathrm{ax}}0`$ Case Even if we switch on $`g_{\mathrm{ax}}0`$, the gap equation does not change. However, the axial current correlator changes in the following way: $`{\displaystyle d^4xe^{iq(xy)}0\left|\text{}TA_\mu ^a(x)A_\nu ^b(y)\right|0}`$ (125) $`=`$ $`\begin{array}{c}{\displaystyle \frac{1}{4}}\delta _{ab}\mathrm{\Pi }_{\mu \nu }(q){\displaystyle \frac{1}{4}}\delta _{ab}\left(g_{\mu \nu }{\displaystyle \frac{q_\nu q_\mu }{q^2}}\right)\mathrm{\Pi }_2(q^2)G_2(q^2)\hfill \\ \text{}{\displaystyle \frac{1}{4}}\delta _{ab}\left({\displaystyle \frac{q_\mu q_\nu }{q^2}}\right)\left(\text{}q^2H(q^2)\mathrm{\Pi }_{5A}(q^2)+G(q^2)\mathrm{\Pi }_1(q^2)\right)\hfill \end{array}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\delta _{ab}\mathrm{\Pi }_{\mu \nu }(q){\displaystyle \frac{1}{4}}\delta _{ab}\left(g_{\mu \nu }{\displaystyle \frac{q_\nu q_\mu }{q^2}}\right){\displaystyle \frac{2ig_{\mathrm{ax}}\mathrm{\Pi }_2(q^2)}{12ig_{\mathrm{ax}}\mathrm{\Pi }_2(q^2)}}{\displaystyle \frac{1}{4}}\delta _{ab}\left({\displaystyle \frac{q_\mu q_\nu }{q^2}}\right){\displaystyle \frac{N(q^2)}{D(q^2)}},`$ where $`H(q^2)`$, $`G_1(q^2)`$, $`G_2(q^2)`$ and $`D(q^2)`$ are defined in Appendix D, and $`N(q^2)`$ is given as follows: $`N(q^2)`$ $`=`$ $`\begin{array}{c}{\displaystyle \frac{1}{4}}q^2(2ig_\pi )\left(\text{}\mathrm{\Pi }_{5A}(q^2)\right)^2\left(\text{}1+2ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\right)\hfill \\ \text{}{\displaystyle \frac{1}{4}}\left(\text{}1+2ig_\pi \mathrm{\Pi }_{55}(q^2)\right)\left(\text{}2ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\right)\hfill \end{array}`$ (131) The explicit expressions for $`m_\pi `$ and $`f_\pi `$ are obtained from: $$0=D(m_\pi ^2),f_\pi =\sqrt{\frac{N(m_\pi ^2)}{im_\pi ^2D^{}(m_\pi ^2)}},$$ (132) where $`D^{}(q^2){\displaystyle \frac{dD(q^2)}{d(q^2)}}`$. ### B.3 The Regularized Bubble Integrals The regularized expressions for the bubble integrals with the sharp Euclidean cut-off are as follows: $`\mathrm{\Pi }_{55}(q^2)`$ $`=`$ $`24i{\displaystyle ^\mathrm{\Lambda }}{\displaystyle \frac{d^4k_E}{(2\pi )^4}}{\displaystyle _0^1}𝑑x{\displaystyle \frac{k_E^2+M^2+q^2x(1x)}{(k_E^2+M^2q^2x(1x))^2}}`$ (133) $`\mathrm{\Pi }_{5A}(q^2)`$ $`=`$ $`24iM{\displaystyle ^\mathrm{\Lambda }}{\displaystyle \frac{d^4k_E}{(2\pi )^4}}{\displaystyle _0^1}𝑑x{\displaystyle \frac{1}{(k_E^2+M^2q^2x(1x))^2}}`$ (134) $`\mathrm{\Pi }_{\mu \nu }(q)`$ $`=`$ $`48i{\displaystyle ^\mathrm{\Lambda }}{\displaystyle \frac{d^4k_E}{(2\pi )^4}}{\displaystyle _0^1}𝑑x{\displaystyle \frac{M^2g_{\mu \nu }x(1x)\left(\text{}q^2g_{\mu \nu }q_\mu q_\nu \right)}{(k_E^2+M^2q^2x(1x))^2}}.`$ (135) These expressions satisfy the following identities: $`\mathrm{\Pi }_{5A}(0)`$ $`=`$ $`2M\mathrm{\Pi }_{55}^{}(0)`$ (136) $`q^\nu \mathrm{\Pi }_{\mu \nu }(q)`$ $`=`$ $`2Mq_\mu \mathrm{\Pi }_{5A}(q^2)`$ (137) $`\mathrm{\Pi }_1(0)`$ $`=`$ $`\mathrm{\Pi }_2(0).`$ (138) We comment here on the expression for $`\mathrm{\Pi }_{\mu \nu }(q)`$. In order to obtain the above expression for $`\mathrm{\Pi }_{\mu \nu }(q)`$, we have to use the following prescription: $$\frac{d^4k}{(2\pi )^4}_0^1𝑑x\frac{2k_\mu k_\nu +g_{\mu \nu }\left(\text{}M^2k^2q^2x(1x)\right)}{(M^2k^2q^2x(1x))^2}0$$ (139) This is due to the following reason. The straight forward application of the sharp Euclidean cut-off leads, rather than to Eq.(137), to the following identity: $$q^\nu \mathrm{\Pi }_{\mu \nu }(q)=q^\nu \mathrm{\Pi }_{\mu \nu }^{(V)}(q)+2Mq_\mu \mathrm{\Pi }_{5A}(q^2),$$ (140) where $`\mathrm{\Pi }_{\mu \nu }^{(V)}(q)`$ is $$\mathrm{\Pi }_{\mu \nu }^{(V)}(q)\delta _{ab}\frac{d^4k}{(2\pi )^4}\text{Tr}\left(\text{}(\gamma _\mu \tau ^a)iS_F(k+q)(\gamma _\nu \tau ^b)iS_F(k)\right).$$ (141) Since the non-vanishing longitudinal part of $`\mathrm{\Pi }_{\mu \nu }^{(V)}(q)`$ is an unphysical cutoff artifact, which spoils the significance of the outputs, the prescription Eq.(139) is often used to suppress it. (See p.178 in ref. .) Note that, if the dimensional regularization is applied, the expression Eq.(139) vanishes identically. Now the analogous subtraction should be performed on the l.h.s. in Eq.(140), otherwise the Gell-Mann-Oakes-Renner relation would be terribly violated in the case $`g_{\mathrm{ax}}0`$. Note that, once we use this prescription, the Gell-Mann-Oakes-Renner violations are less than 1% for $`m_\pi 140`$ MeV. ## Appendix C Matrix Element in terms of the 3qBS/Faddeev Amplitudes in the Rest Frame In principle, we can use any Lorentz frames to calculate the Lorentz invariant form factors of a matrix element. But in practice, it is convenient to make use of the rest frame, because this frame is often used in the calculation of the mass and the 3qBS amplitudes. The aim of this appendix is to give expressions for the matrix element in terms of the 3qBS /Faddeev amplitudes in the rest frame by using the Lorentz covariance. Unless the opposite is explicitly indicated, the external fields are understood to be absent, i.e., $`v=a=0`$, $`m=m_0`$. To make the Lorentz transformation of the spinor simpler, we select the final momentum $`P=(m_N,\stackrel{}{0})`$ and the initial momentum $`L^\mu =(\mathrm{\Lambda }^1)_{}^{\mu }{}_{\nu }{}^{}P^\nu `$, where $`\mathrm{\Lambda }_{}^{\mu }{}_{\nu }{}^{}=(e^\omega )_{}^{\mu }{}_{\nu }{}^{}`$ represents the boost. We introduce the Fourier transforms of the 3qBS amplitudes and of $`O_\mu ^b\left[{\displaystyle \frac{i\delta K^{[e]}}{\delta a_\mu ^b(0)}}\right]^{[0]}`$: $`\mathrm{\Psi }[N_\beta ^c(L)](l,l^{})(2\pi )^4\delta ^{(4)}(L\overline{L})`$ $`=`$ $`{\displaystyle d^4y_1d^4y_2d^4y_3e^{i{\scriptscriptstyle l_iy_i}}\mathrm{\Psi }[N_\beta ^c(L)](y_1,y_2,y_3)}`$ $`\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(P)](p,p^{})(2\pi )^4\delta ^{(4)}(\overline{P}P)`$ $`=`$ $`{\displaystyle d^4x_1d^4x_2d^4x_3\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(P)](x_1,x_2,x_3)e^{i{\scriptscriptstyle p_ix_i}}}`$ $`O_\mu ^b(\overline{P},p,p^{};\overline{L},l,l^{})`$ $`=`$ $`{\displaystyle d^4x_1d^4x_2d^4x_3d^4y_1d^4y_2d^4y_3e^{i{\scriptscriptstyle p_ix_i}}O_\mu ^b(x_1,x_2,x_3;y_1,y_2,y_3)e^{i{\scriptscriptstyle l_iy_i}}}`$ $`O_\mu ^b(P,L;p,p^{},l,l^{})`$ $`=`$ $`{\displaystyle \frac{d^4\overline{P}}{(2\pi )^4}\frac{d^4\overline{L}}{(2\pi )^4}(2\pi )^4\delta ^{(4)}(\overline{P}P)O_\mu ^b(\overline{P},p,p^{};\overline{L},l,l^{})(2\pi )^4\delta ^{(4)}(L\overline{L})}`$ $`\overline{P},\overline{L}`$ are the total momenta and $`p,p^{},l,l^{}`$ are the relative momenta, which are defined though the following relations: $$\begin{array}{ccc}\overline{P}=p_i\hfill & p=\eta p_3(1\eta )(p_1+p_2),\hfill & p^{}=\eta ^{}p_1(1\eta ^{})p_2\hfill \\ \text{}\overline{L}=l_i\hfill & l=\eta l_3(1\eta )(l_1+l_2),\hfill & l^{}=\eta ^{}l_1(1\eta ^{})l_2,\hfill \end{array}$$ (143) where $`0<\eta ,\eta ^{}<1`$ are arbitrary real numbers. We note that, unlike the nonrelativistic approaches, the choice of $`\eta `$ and $`\eta ^{}`$ is almost completely arbitrary in the relativistic quantum field theory<sup>21</sup><sup>21</sup>21For complete expositions, see .. The delta functions in the first two relations in Eq.(C) are due to the translational invariance. It is important to note that the Jacobians associated with the variable change, i.e., $`(p_1,p_2,p_3)(\overline{P},p,p^{})`$ and $`(l_1,l_2,l_3)(\overline{L},l,l^{})`$ are 1, i.e., $$\frac{d^4p_1}{(2\pi )^4}\frac{d^4p_2}{(2\pi )^4}\frac{d^4p_3}{(2\pi )^4}=\frac{d^4\overline{P}}{(2\pi )^4}\frac{d^4p}{(2\pi )^4}\frac{d^4p^{}}{(2\pi )^4},\frac{d^4l_1}{(2\pi )^4}\frac{d^4l_2}{(2\pi )^4}\frac{d^4l_3}{(2\pi )^4}=\frac{d^4\overline{L}}{(2\pi )^4}\frac{d^4l}{(2\pi )^4}\frac{d^4l^{}}{(2\pi )^4}.$$ (144) Now the matrix element reduces to the following expression: $`N_\alpha ^a(P)\left|\text{}A_\mu ^b(0)\right|N_\beta ^c(L)\left(\stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(P)]O_\mu ^b\mathrm{\Psi }[N_\beta ^c(L)]\right)`$ (145) $`=`$ $`\begin{array}{c}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{d^4p^{}}{(2\pi )^4}\frac{d^4l}{(2\pi )^4}\frac{d^4l^{}}{(2\pi )^4}}\hfill \\ \text{}\times \stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(P)](p,p^{})O_\mu ^b(P,L;p,p^{},l,l^{})\mathrm{\Psi }[N_\beta ^c(L)](l,l^{})\hfill \end{array}`$ (148) $`=`$ $`\begin{array}{c}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{d^4p^{}}{(2\pi )^4}\frac{d^4l}{(2\pi )^4}\frac{d^4l^{}}{(2\pi )^4}}\hfill \\ \text{}\times \stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(P)](p,p^{})O_\mu ^b(P,L;p,p^{},l,l^{})\widehat{S}(\mathrm{\Lambda }^1)^3\mathrm{\Psi }[N_\beta ^c(P)](\mathrm{\Lambda }l,\mathrm{\Lambda }l^{})\hfill \end{array}`$ (151) $`=`$ $`\begin{array}{c}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{d^4p^{}}{(2\pi )^4}\frac{d^4l}{(2\pi )^4}\frac{d^4l^{}}{(2\pi )^4}}\hfill \\ \text{}\times \stackrel{~}{\overline{\mathrm{\Psi }}}[N_\alpha ^a(P)](p,p^{})O_\mu ^b(P,L;p,p^{},\mathrm{\Lambda }^1l,\mathrm{\Lambda }^1l^{})\widehat{S}(\mathrm{\Lambda }^1)^3\mathrm{\Psi }[N_\beta ^c(P)](l,l^{}),\hfill \end{array}`$ (154) where the second equality follows from the following Lorentz transformation property of the 3qBS amplitude of the rest frame: $`0\left|\text{}T\psi (x_1)\psi (x_2)\psi (x_3)\right|N_\beta ^c(\mathrm{\Lambda }^1P)`$ $`=`$ $`0\left|\text{}\left[T\psi (x_1)\psi (x_2)\psi (x_3)\right]\stackrel{ˇ}{S}(\mathrm{\Lambda }^1)\right|N_\beta ^c(P)`$ $`=`$ $`\widehat{S}(\mathrm{\Lambda }^1)^30\left|\text{}T\psi (\mathrm{\Lambda }x_1)\psi (\mathrm{\Lambda }x_2)\psi (\mathrm{\Lambda }x_3)\right|N_\beta ^c(P),`$ where $`(\mathrm{\Lambda }^1)_{\mu }^{}{}_{}{}^{\nu }=(e^\lambda )_{\mu }^{}{}_{}{}^{\nu }`$ is the pure boost matrix, $`\stackrel{ˇ}{S}(\mathrm{\Lambda }^1)`$ is the boost operator in the Fock space, $`\widehat{S}(\mathrm{\Lambda }^1)=e^{S(\lambda )}`$, ($`S(\lambda )={\displaystyle \frac{i}{4}}\sigma _{\mu \nu }(\lambda )^{\mu \nu }`$) the boost matrix in the Dirac bispinor space, and $`\widehat{S}(\mathrm{\Lambda })^3=\widehat{S}(\mathrm{\Lambda })\widehat{S}(\mathrm{\Lambda })\widehat{S}(\mathrm{\Lambda })`$. In the last line of Eq.(145), with the aid of the Lorentz invariance of the Jacobian, we rotate the integration variables $`l,l^{}`$ by the boost $`\mathrm{\Lambda }`$, i.e., $`l\mathrm{\Lambda }^1l,l^{}\mathrm{\Lambda }^1l^{}`$. The problem thus reduces to the “matrix element” calculation of $`O_\mu ^b(P,\mathrm{\Lambda }^1P;p,p^{},\mathrm{\Lambda }^1l,\mathrm{\Lambda }^1l^{})\widehat{S}(\mathrm{\Lambda }^1)^3`$ between two 3qBS amplitudes in the rest frame. By using the homogeneous Faddeev equation Eq.(104), the Faddeev amplitude is expressed by the 3qBS amplitude as follows: $$\psi [N_\alpha ^a(\stackrel{}{P})]=K_3\mathrm{\Psi }[N_\alpha ^a(\stackrel{}{P})].$$ (156) Hence, the Lorentz transformation property of the 3qBS amplitude implies the following Lorentz transformation property of the Faddeev amplitude: $$\psi [N_\alpha ^a(\mathrm{\Lambda }^1P)](x_1,x_2,x_3)=\widehat{S}(\mathrm{\Lambda }^1)^3\psi [N_\alpha ^a(P)](\mathrm{\Lambda }x_1,\mathrm{\Lambda }x_2,\mathrm{\Lambda }x_3).$$ (157) By repeating almost the same arguments, we are left with the following expression of the matrix element in terms of the Faddeev amplitudes in the rest frame: $`N_\alpha ^a(P)\left|\text{}A_\mu ^b(x=0)\right|N_\beta ^c(L)`$ (158) $`=`$ $`\begin{array}{c}{\displaystyle \frac{d^4p}{(2\pi )^4}\frac{d^4p^{}}{(2\pi )^4}\frac{d^4l}{(2\pi )^4}\frac{d^4l^{}}{(2\pi )^4}}\hfill \\ \text{}\times \overline{\psi }[N_\alpha ^a(P)](p,p^{})O_{F;\mu }^b(P,L;p,p^{},\mathrm{\Lambda }^1,\mathrm{\Lambda }^1l^{})\widehat{S}(\mathrm{\Lambda }^1)^3\psi [N_\beta ^c(P)](l,l^{}),\hfill \end{array}`$ (161) where the Fourier transforms $`\psi [N_\beta ^c(P)](p,p^{})`$, $`\overline{\psi }[N_\alpha ^a(P)](p,p^{})`$ are defined from $`\psi [N_\beta ^c(P)](x_1,x_2,x_3)`$, $`\overline{\psi }[N_\alpha ^a(P)](x_1,x_2,x_3)`$ similar to Eq.(C), and $`O_{F;\mu }^b(P,L;p,p^{},l,l^{})`$ is defined from: $$O_{F;\mu }^b(x_1,x_2,x_3;y_1,y_2,y_3)\left[\text{}\frac{i\delta K_F^{[e]}(x_1,x_2,x_3;y_1,y_2,y_3)}{\delta a_\mu ^b(x=0)}\right]^{[0]}.$$ (162) ## Appendix D Non-vanishing $`g_{\mathrm{ax}}`$ ### D.1 General Formalism The aim of this appendix is to consider how to extract $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ in the case $`g_{\mathrm{ax}}0`$ providing an analytical expression of the GT violation $`\mathrm{\Delta }_G(m_0)`$. Instead of repeating the argument similar to the one given in Section 5.1, which would become quite lengthy, we derive the expression of the form factors in a different manner, which would be easier for the readers to understand. We first have to consider functional derivatives of the constituent quark propagator, i.e., $`\delta S_F/\delta a_\mu ^b`$ and $`\delta S_F/\delta p^b`$, which are obtained as the solutions to Eq.(42) and Eq.(82). We parameterize $`\delta \mathrm{\Sigma }/\delta a_\mu ^b`$ and $`\delta \mathrm{\Sigma }/\delta p^b`$ as follows: $`{\displaystyle d^4z^{}e^{iq(z^{}z)}\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta a_\mu ^b(z)}\right]^{[0]}}`$ $`=`$ $`H(q^2)\left(q^\mu \gamma _5{\displaystyle \frac{\tau ^b}{2}}\right)+G_1(q^2)\left({\displaystyle \frac{q^\mu \overline{)}q}{q^2}}\gamma _5{\displaystyle \frac{\tau ^b}{2}}\right)+G_2(q^2)\left(\gamma ^\mu {\displaystyle \frac{q^\mu \overline{)}q}{q^2}}\right)\gamma _5{\displaystyle \frac{\tau ^b}{2}}`$ $`{\displaystyle d^4z^{}e^{iq(z^{}z)}\left[\frac{\delta \mathrm{\Sigma }^{[e]}(z^{})}{\delta p^b(z)}\right]^{[0]}}`$ $`=`$ $`I_1(q^2)\left(i\gamma _5{\displaystyle \frac{\tau _b}{2}}\right)+I_2(q^2)\left(i\overline{)}q\gamma _5{\displaystyle \frac{\tau _b}{2}}\right).`$ $`H(q^2)`$, $`G_1(q^2)`$, $`G_2(q^2)`$, $`I_1(q^2)`$ and $`I_2(q^2)`$ satisfy the following coupled equations: $`G_2(q^2)`$ $`=`$ $`2ig_{\mathrm{ax}}\mathrm{\Pi }_2(q^2)+2ig_{\mathrm{ax}}\mathrm{\Pi }_2(q^2)G_2(q^2)`$ (164) $`\left(\begin{array}{c}H(q^2)\hfill \\ G_1(q^2)\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{c}2ig_\pi \mathrm{\Pi }_{5A}(q^2)\hfill \\ 2ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\hfill \end{array}\right)+\left(\begin{array}{cc}2ig_\pi \mathrm{\Pi }_{55}(q^2)\hfill & 2ig_\pi \mathrm{\Pi }_{5A}(q^2)\hfill \\ q^22ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)\hfill & 2ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\hfill \end{array}\right)\left(\begin{array}{c}H(q^2)\hfill \\ G_1(q^2)\hfill \end{array}\right)`$ (173) $`\left(\begin{array}{c}I_1(q^2)\hfill \\ I_2(q^2)\hfill \end{array}\right)`$ $`=`$ $`\left(\begin{array}{c}2ig_\pi \mathrm{\Pi }_{55}(q^2)\hfill \\ 2ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)\hfill \end{array}\right)+\left(\begin{array}{cc}2ig_\pi \mathrm{\Pi }_{55}(q^2)\hfill & q^2(2ig_\pi )\mathrm{\Pi }_{5A}(q^2)\hfill \\ 2ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)\hfill & 2ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\hfill \end{array}\right)\left(\begin{array}{c}I_1(q^2)\hfill \\ I_2(q^2)\hfill \end{array}\right),`$ (182) with the following solutions: $`H(q^2)`$ $`=`$ $`{\displaystyle \frac{2ig_\pi \mathrm{\Pi }_{5A}(q^2)}{D(q^2)}}`$ (183) $`G_1(q^2)`$ $`=`$ $`{\displaystyle \frac{q^2\left(\text{}2ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)\right)\left(\text{}2ig_\pi \mathrm{\Pi }_{5A}(q^2)\right)+\left(\text{}1+2ig_\pi \mathrm{\Pi }_{55}(q^2)\right)\left(\text{}2ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\right)}{D(q^2)}}`$ $`G_2(q^2)`$ $`=`$ $`{\displaystyle \frac{2ig_{\mathrm{ax}}\mathrm{\Pi }_2(q^2)}{12ig_{\mathrm{ax}}\mathrm{\Pi }_2(q^2)}}`$ $`I_1(q^2)`$ $`=`$ $`{\displaystyle \frac{\left(\text{}12ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\right)\left(\text{}2ig_\pi \mathrm{\Pi }_{55}(q^2)\right)+q^2\left(\text{}2ig_\pi \mathrm{\Pi }_{5A}(q^2)\right)\left(\text{}2ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)\right)}{D(q^2)}}`$ $`I_2(q^2)`$ $`=`$ $`{\displaystyle \frac{2ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)}{D(q^2)}}`$ $`D(q^2)`$ $`=`$ $`\left(\text{}1+2ig_\pi \mathrm{\Pi }_{55}(q^2)\right)\left(\text{}12ig_{\mathrm{ax}}\mathrm{\Pi }_1(q^2)\right)q^2\left(\text{}2ig_\pi \mathrm{\Pi }_{5A}(q^2)\right)\left(\text{}2ig_{\mathrm{ax}}\mathrm{\Pi }_{5A}(q^2)\right).`$ We define “baryonic parts” $`B_P(q^2)`$ and $`B_A(q^2)`$ of the form factors through Fig.17. Now the matrix elements are expressed as follows: $`N(P)\left|\text{}A_\mu ^b(x=0)\right|N(L)=\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}\left(\text{}g_A(q^2)(\gamma _\mu \gamma _5)+h_A(q^2)(q_\mu \gamma _5)\right)u(L)`$ (184) $`=`$ $`\begin{array}{c}\left(\text{}\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}\gamma _\mu \gamma _5u(L)\right)B_A(q^2)\left(\text{}1+G_2(q^2)\right)\hfill \\ \text{}+\left(\text{}\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}\overline{)}q\gamma _5u(L)\right)B_A(q^2)q_\mu {\displaystyle \frac{G_1(q^2)G_2(q^2)}{q^2}}\hfill \\ \text{}+\left(\text{}\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}\gamma _5u(L)\right)B_P(q^2)q_\mu H(q^2)\hfill \end{array}`$ (188) $`N(P)\left|\text{}i\overline{\psi }(0)\gamma _5{\displaystyle \frac{\tau ^b}{2}}\psi (0)\right|N(L)=\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}\left(\text{}i_A(q^2)(i\gamma _5)\right)u(L)`$ $`=`$ $`\left(\text{}\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}i\gamma _5u(L)\right)B_P(q^2)\left(\text{}1+I_1(q^2)\right)+\left(\text{}\overline{u}(P){\displaystyle \frac{\tau ^b}{2}}i\overline{)}q\gamma _5u(L)\right)B_A(q^2)I_2(q^2),`$ where the isospin and the helicity indices of the initial and the final nucleons are suppressed for simplicity. The form factors are expressed as $`g_A(q^2)`$ $`=`$ $`B_A(q^2)\left(\text{}1+G_2(q^2)\right)`$ (190) $`h_A(q^2)`$ $`=`$ $`B_P(q^2)H(q^2)+2m_NB_A(q^2){\displaystyle \frac{G_1(q^2)G_2(q^2)}{q^2}}`$ $`i_A(q^2)`$ $`=`$ $`B_P(q^2)\left(\text{}1+I_1(q^2)\right)+2m_NB_A(q^2)I_2(q^2).`$ $`\stackrel{~}{g}_{\pi NN}`$ is extracted according to Eq.(86) and $`g_A`$ by $`g_A=g_A(q^2=0)`$. Note that, due to Eq.(138), $`G_1(q^2)G_2(q^2)`$ is proportional to $`q^2`$, which cancel $`q^2`$ in the denominator in the expression of $`h_A(q^2)`$. By neglecting the small cut-off artifact, the PCAC relation Eq.(34) leads to the relation: $$2m_Ng_A(q^2)+q^2h_A(q^2)=2m_0i_A(q^2),$$ (191) which provides the following relation between $`B_P(q^2)`$ and $`B_A(q^2)`$: $$2m_NB_A(q^2)\left(\text{}1+G_1(q^2)2m_0I_2(q^2)\right)+B_P(q^2)\left(\text{}q^2H(q^2)2m_0(1+I_1(q^2))\right)=0.$$ (192) By using Eq.(183), Eq.(190) and Eq.(192), we are left with the following analytical expression of $`\mathrm{\Delta }_\mathrm{G}(m_0)`$: $`\mathrm{\Delta }_\mathrm{G}(m_0)`$ $`=`$ $`{\displaystyle \frac{f_\pi \stackrel{~}{g}_{\pi NN}}{m_Ng_A}}`$ $`=`$ $`{\displaystyle \frac{m_\pi ^2}{2m_0}}{\displaystyle \frac{2ig_\pi \mathrm{\Pi }_{5A}(0)}{12ig_{\mathrm{ax}}\mathrm{\Pi }_1(0)}}+m_\pi ^2{\displaystyle \frac{2ig_{\mathrm{ax}}}{12ig_A\mathrm{\Pi }_1(0)}}\left(\text{}\mathrm{\Pi }_1^{}(0)\mathrm{\Pi }_2^{}(0)\right),`$ where $`\mathrm{\Pi }_1^{}(q^2){\displaystyle \frac{d\mathrm{\Pi }_1(q^2)}{d(q^2)}}`$, $`\mathrm{\Pi }_2^{}(q^2){\displaystyle \frac{d\mathrm{\Pi }_2(q^2)}{d(q^2)}}`$. All the baryonic quantities disappear from $`\mathrm{\Delta }(m_0)`$. Note that, once the mean-field approximation for the vacuum is adopted, this expression is valid even beyond the ladder truncation scheme of the 3qBS kernel, as far as the kernel satisfies our criterion of PCAC relation<sup>22</sup><sup>22</sup>22Although the baryonic parts $`B_P(q^2)`$ and $`B_A(q^2)`$ change, they still satisfy Eq.(192). Hence Eq.(D.1) remains to be valid.. This, in particular, implies that, to improve $`\mathrm{\Delta }_\mathrm{G}(m_0)`$, we have to improve the vacuum or to evaluate the on-shell $`g_{\pi NN}`$. Therefore, the easiest way to improve the result of $`\mathrm{\Delta }_\mathrm{G}(m_0)`$ in Section 5.4 is to improve the vacuum by including the effect of $`g_{\mathrm{ax}}0`$, since it is not so easy to treat the vacuum beyond the meanfield approximation. ### D.2 The Numerical Results We first have to explain the choice of the parameters. There are five parameters: the cutoff $`\mathrm{\Lambda }`$ (Euclidean sharp cutoff), the current quark mass $`m_0`$, the effective coupling constant in the pionic channel $`g_\pi `$, the effective coupling constant in the iso-vector axial-vector mesonic channel $`g_{\mathrm{ax}}`$ and the effective coupling constant in the $`qq`$ scalar diquark channel $`g_{\mathrm{sd}}`$. We treat $`g_{\mathrm{ax}}`$ as a free parameter, and fix the first three parameters ($`\mathrm{\Lambda },m_0,g_\pi `$) by solving the gap equation Eq.(118) for the constituent quark mass $`M`$, and Eq.(132) for the pion mass $`m_\pi `$ and the pion decay constant $`f_\pi `$ imposing the following three conditions: (1) $`m_\pi =140`$ MeV, (2) $`f_\pi =93`$ MeV, (3) $`M=400`$ MeV. We fix $`g_{\mathrm{sd}}`$ by requiring that $`m_N=940`$ MeV. We give our numerical results for the three cases (1) $`g_{\mathrm{ax}}/g_\pi =0`$, (2) $`g_{\mathrm{ax}}/g_\pi =0.25`$, (3) $`g_{\mathrm{ax}}/g_\pi =0.5`$. The explicit values of the parameters are listed as follows: 1. $`g_{\mathrm{ax}}/g_\pi =0`$: $`\mathrm{\Lambda }=739`$ MeV, $`m_0=8.99`$ MeV, $`g_\pi =10.4`$ GeV<sup>-2</sup>, $`g_\mathrm{s}/g_\pi =0.66`$. 2. $`g_{\mathrm{ax}}/g_\pi =0.25`$: $`\mathrm{\Lambda }=812`$ MeV, $`m_0=6.99`$ MeV, $`g_\pi =8.13`$ GeV<sup>-2</sup>, $`g_\mathrm{s}/g_\pi =0.694`$. 3. $`g_{\mathrm{ax}}/g_\pi =0.5`$: $`\mathrm{\Lambda }=874`$ MeV, $`m_0=5.75`$ MeV, $`g_\pi =6.71`$ GeV<sup>-2</sup>, $`g_\mathrm{s}/g_\pi =0.72`$. Once these parameters are fixed, we consider the chiral limit by taking $`m_00`$ keeping $`\mathrm{\Lambda }`$, $`g_\pi `$, $`g_{\mathrm{ax}}`$ and $`g_{\mathrm{sd}}`$ fixed. Eq.(118) provides the $`m_0`$ dependences of $`M=M(m_0)`$. We confine ourselves to non-negative values of $`g_{\mathrm{ax}}/g_\pi `$ because of the following reasons: (1) our examples of the interaction Lagrangians give non-negative values, i.e., $`g_{\mathrm{ax}}/g_\pi =0`$ in the original NJL type, and $`g_{\mathrm{ax}}/g_\pi =0.5`$ in the color current type. (2) As we shall see below, $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ increase with decreasing $`g_{\mathrm{ax}}/g_\pi `$. If $`g_{\mathrm{ax}}<0`$, we cannot adjust either $`g_A`$ nor $`\stackrel{~}{g}_{\pi NN}`$. (3) The increase of $`g_A`$ with decreasing $`g_{\mathrm{ax}}`$ is due to the fact that the negative $`g_{\mathrm{ax}}/g_\pi `$ works as an attraction in the transversal iso-vector axial-vector mesonic channel ($`a_1`$ channel), which makes $`G_2(q^2=0)`$ to grow up, leading to the rapid increase of $`g_A=g_A(q^2=0)`$. (cf. Eq.(190)) This “anti-screening” of $`g_A`$ does not seem to be reasonable. Now we present our numerical results. In Fig.18, we plot $`g_{\pi NN}`$ against the current quark mass $`m_0`$ for these three cases of $`g_{\mathrm{ax}}`$. It is seen that $`\stackrel{~}{g}_{\pi NN}`$ decreases with increasing $`g_{\mathrm{ax}}/g_\pi `$, and increases with increasing $`m_0`$. The crosses in the figure are used to indicate $`m_0`$ which correspond to $`m_\pi =140`$ MeV. We use the diamonds to indicate $`m_0`$ which correspond to $`m_\pi =2\times 140`$ MeV, which is used to indicate the validity of the single pole dominance approximation of $`\stackrel{~}{g}_{\pi NN}`$. Note that the distance between $`0`$ and $`4M^2`$, i.e., the $`q\overline{q}`$ cut, is still $`10`$ times larger than $`m_\pi ^2`$, as long as $`m_\pi 2\times 140`$ MeV. In Fig.19, we plot $`g_A`$ against $`m_0`$ for the three cases of $`g_{\mathrm{ax}}`$. We see that $`g_A`$ decreases with increasing $`g_{\mathrm{ax}}/g_\pi `$ and increases with increasing $`m_0`$. Note that, if $`g_{\mathrm{ax}}/g_\pi =0.1`$, then $`g_A=1.23`$ and $`\stackrel{~}{g}_{\pi NN}=12.5`$. So the best fit of $`\stackrel{~}{g}_{\pi NN}`$ and $`g_A`$ could be obtained in the region $`0g_{\mathrm{ax}}/g_\pi <0.1`$, which, however, would depend on the quantity which we prefer to adjust. From this point of view, $`g_{\mathrm{ax}}=0`$ is actually a rather good choice. In this case, $`\stackrel{~}{g}_{\pi NN}=13.5`$ is very reasonable and $`g_A=1.33`$ is still close to the experimental value $`g_A^{(\mathrm{exp})}=1.26`$. Note that, due to the chiral symmetry, positive $`g_{\mathrm{ax}}/g_\pi `$ implies an attractive interaction in the iso-vector vector mesonic channel ($`\rho `$ meson). In order to describe $`\rho `$ meson in the NJL model, however, we need stronger $`g_{\mathrm{ax}}`$ , but, stronger $`g_{\mathrm{ax}}`$ leads to smaller $`g_A`$ and $`\stackrel{~}{g}_{\pi NN}`$ in our calculation. (The situation could be improved by including $`qq`$ interactions in the axial-vector diquark channel, which enhance $`g_{\pi NN}`$ . In this case, one should also include the $`qq`$ interaction in the vector diquark channel for the Goldberger-Treiman and the PCAC relation.) In Fig.20, we plot the numerical $`\mathrm{\Delta }_\mathrm{G}`$ together with the analytic $`\mathrm{\Delta }_\mathrm{G}`$ (Eq.(D.1)) against $`m_0`$. It is seen that they are monotonically decreasing functions of $`m_0`$ independent of $`g_{\mathrm{ax}}/g_\pi `$. We see that the discrepancy between the analytic and the numeric $`\mathrm{\Delta }_\mathrm{G}`$ is within 3 %. Recall that, when deriving Eq.(D.1), we neglected the small cut-off artifact. Hence this discrepancy is solely due to the cut-off artifact. Note that the analytic $`\mathrm{\Delta }_\mathrm{G}`$ approaches $`1`$ as $`m_00`$ in Fig.20. In fact, based on Eq.(136), Eq.(137), Eq.(138) and the Gell-Mann-Oakes-Renner relation, it is possible to prove that the analytic $`\mathrm{\Delta }_\mathrm{G}`$ approaches to $`1`$ as $`m_00`$. To make $`\mathrm{\Delta }_\mathrm{G}`$ an increasing function of $`m_0`$, it is necessary to go beyond the validity of the analytic $`\mathrm{\Delta }_\mathrm{G}`$ (Eq.(D.1)). This implies that all we can do is either to improve the gap equation for the vacuum beyond the meanfield approximation or to evaluate the on-shell $`g_{\pi NN}`$. In Fig.21, we plot the PCAC violation $`\mathrm{\Delta }_\mathrm{P}`$ against $`m_0`$ for the three cases of $`g_{\mathrm{ax}}/g_\pi `$. The deviation of $`\mathrm{\Delta }_P`$ from $`1`$ is solely due to the cutoff artifact. We see that the PCAC violations are only within 3%, which suggests the practical validity of our results. ## Figure Captions 1. The gauge transformation property of $`S_F^{[e]}`$ is depicted. A single line represents a propagator of the constituent quark in the presence of the external fields. The gauge transformation reduces to the multiplication of the phase factors $`\mathrm{\Omega }(x)`$ and $`\mathrm{\Omega }(y)`$. 2. The functional derivative of the propagator, i.e., $`\left[\text{}{\displaystyle \frac{i\delta S_F^{[e]}}{\delta a_\mu ^b(z)}}\right]^{[0]}`$ is depicted. A single line is a constituent quark propagator. Dashed lines are used to indicate the momentum transfer $`q`$. Fig.2.(a) corresponds to $`g_{\mathrm{ax}}=0`$ case, and Fig.2.(b) corresponds to $`g_{\mathrm{ax}}0`$ case. 3. (a) The 3qBS kernel in the ladder truncation scheme is depicted. A single line represents a quark propagator. A slash indicates an amputation, i.e., removement of the quark leg. (b) The diagram used to evaluate the nucleon matrix element for 3qBS kernel of the type Fig.3.(a) is depicted. “$``$” is understood in the sense of Fig.2. Diagrams of the same topologies are omitted. 4. (a) The $`q\overline{q}`$ exchange improved 3qBS kernel is depicted. A single line represents a quark propagator. A slash indicates an amputation. A wavy line represents a ladder sum of $`q\overline{q}`$ bubbles as is depicted in the second line. Diagrams of the same topologies are omitted. (b) The diagrams with which to obtain the nucleon matrix element for 3qBS kernel of the type Fig.4.(a) are depicted. The second term is often referred to as the “meson exchange current” contribution. The diagrams of the same topologies are omitted. 5. The proof that the ladder truncated 3qBS kernel gives rise to the PCAC relation correctly is diagrammatically explained. In the first step, the local gauge transformation $`\mathrm{\Omega }(x)=e^{i\gamma _5\omega (x)}`$ is applied. Only the constituent quark propagators transform. In the next step, Eq.(54) is used. The last step is due to the fact that amputated propagators are delta functions. It is seen that the sufficient condition is satisfied. 6. The proof that the $`q\overline{q}`$ exchange improved 3qBS kernel gives rise to the PCAC relation correctly is diagrammatically explained. In the first step, we applied the gauge transformation. In the second step, Eq.(49) is used. Note that all the phase factors, which appear at the internal vertices, cancel themselves. The last step is due to the fact that the amputated propagators are delta functions. It is seen that the sufficient condition is satisfied. 7. (a) The Faddeev equation Eq.(96) is depicted. A single line is a constituent quark propagator. A double line is a t-matrix in the $`qq`$ diquark channel. (For precise meaning, see Fig.7.(b).) (b) The diagramatic expression of the two-quark resolvent $`{\displaystyle \frac{K_3}{1K_3}}`$ is depicted. A single line is a constituent quark propagator. A slash indicates an amputation. 8. The diagrams which contribute to the matrix element $`N|A_\mu ^b(x)|N`$ are depicted. A blob with “N” which is followed by a triple line is a quark-diquark Faddeev amplitude of the nucleon. A single line is a constituent quark propagator, and a double line is a t-matrix in the $`qq`$ scalar diquark channel. A dashed line indicates a momentum transfer $`q_\mu `$. Fig.8.(a) is referred to as the “quark current” contribution, and Fig.8.(b) the “exchange current” contribution. Fig.8.(c) is referred to as the “diquark current” contribution. If the $`qq`$ interaction is truncated to the scalar diquark channel, Fig.8.(c) does not contribute to $`g_A`$, due to the iso-scalar nature of the scalar diquark. Fig.8.(a’) is equivalent to Fig.8.(a), which is obtained by once iterating the Faddeev amplitude in the finial state by using the homogeneous Faddeev equation. By using Eq.(158) and $`\gamma ^\mu (\mathrm{\Lambda }p)_\mu =S(\mathrm{\Lambda })\overline{)}pS(\mathrm{\Lambda }^1)`$, Fig.8.(a’) and Fig.8.(b) reduce to the four diagrams depicted in Fig.8.(d). 9. Diagrams which contribute to $`g_A`$ and $`g_{\pi NN}`$ are depicted. These eight diagrams are equivalent to four diagrams in Fig.8.(d) in the limit $`q0`$, which is indicated by “$`q0`$”. The explicit spin-parity projection shows that the first two contribute to $`g_A`$, and the others to $`g_{\pi NN}`$. 10. The nucleon mass $`m_N`$ (for $`g_{\mathrm{sd}}/g_{pi}=0.66`$) (solid line), the quark-diquark threshold $`M+m_{\mathrm{sd}}`$ ($`g_{\mathrm{sd}}/g_\pi =0.66`$ case) (dotted line), the pion mass $`m_\pi `$ (dashed line), the pion decay constant $`f_\pi `$ (dot-dashed line), the constituent quark mass $`M`$ (dot-dot-dashed line) are plotted against the current quark mass $`m_0`$. The points which corresponds to $`m_\pi =140`$ MeV are indicated by the vertical dotted line. 11. The nucleon mass is plotted against $`g_{\mathrm{sd}}/g_\pi `$ for the two cases (1) the chiral limit (solid line), (2) off the chiral limit \[$`m_\pi =140`$ MeV\] (dashed line). The vertical dotted line indicates $`g_{\mathrm{sd}}/g_\pi =0.66`$, where $`m_N`$ becomes $`940`$ MeV for the case $`m_\pi =140`$ MeV. 12. $`g_{\pi NN}`$ is plotted against $`g_{\mathrm{sd}}/g_\pi `$ for the two cases (1) the chiral limit (solid line), (2) off the chiral limit \[$`m_\pi =140`$ MeV\] (dashed line). The vertical dotted line indicates $`g_{\mathrm{sd}}/g_\pi =0.66`$, where $`m_N`$ becomes $`940`$ MeV for the case $`m_\pi =140`$ MeV. 13. The iso-vector $`g_A`$ is plotted against $`g_{\mathrm{sd}}/g_\pi `$ for the two cases (1) the chiral limit (solid line), (2) off the chiral limit \[$`m_\pi =140`$ MeV\] (dashed line). The vertical dotted line indicates $`g_{\mathrm{sd}}/g_\pi =0.66`$, where $`m_N`$ becomes $`940`$ MeV for the case $`m_\pi =140`$ MeV. 14. The violation of the GT relation $`\mathrm{\Delta }_\mathrm{G}`$ is plotted against $`g_{\mathrm{sd}}/g_\pi `$ for the two cases (1) the chiral limit (solid line), (2) off the chiral limit \[$`m_\pi =140`$ MeV\] (dotted line). The dashed line is the plot of the violation of the PCAC relation $`\mathrm{\Delta }_\mathrm{P}`$ off the chiral limit. The vertical dotted line indicates $`g_{\mathrm{sd}}/g_\pi =0.66`$, where $`m_N`$ becomes $`940`$ MeV for the case $`m_\pi =140`$ MeV. 15. The violation of the GT relation $`\mathrm{\Delta }_\mathrm{G}`$ (solid line) and the violation of the PCAC relation $`\mathrm{\Delta }_\mathrm{P}`$ (dashed line) are plotted against the current quark mass $`m_0`$ for the case $`g_{\mathrm{sd}}/g_\pi =0.66`$. The vertical dotted line indicates the current quark mass $`m_0`$ which corresponds to $`m_\pi =140`$ MeV. 16. The interaction Lagrangian $`_I`$ is depicted in Fig.16.(a). The diagramatic interpretations of this interaction in the $`q\overline{q}`$ channel are classified into the two types: the direct channel \[Fig.16.(b)\] and the exchange channel (Fig.16.(c)). Fig.16.(b) is obtained from Fig.16.(a) by using the Fierz identity. Another Fierz identity leads to the diagramatic representation of the interaction in the $`qq`$ channel which is depicted in Fig.16.(d). 17. The baryonic parts $`B_P(q^2)`$ and $`B_A(q^2)`$ of the form factors are defined diagrammatically. A blob with “N” which is followed by a triple line is a quark-diquark Faddeev amplitude of the nucleon. A single line is the constituent quark propagator, and a double line is a t-matrix in the $`qq`$ scalar diquark channel. A dashed line indicates a momentum transfer $`q_\mu `$. 18. $`g_{\pi NN}`$ is plotted against the current quark mass $`m_0`$ for the three cases (1) $`g_{\mathrm{ax}}/g_\pi =0`$ (solid line), (2) $`g_{\mathrm{ax}}/g_\pi =0.25`$ (dotted line), (3) $`g_{\mathrm{ax}}/g_\pi =0.5`$ (dashed line). The crosses indicate the points which correspond to $`m_\pi =140`$ MeV, and the diamonds indicate the points which correspond to $`m_\pi =280`$ MeV. 19. $`g_A`$ is plotted against the current quark mass $`m_0`$ for the three cases (1) $`g_{\mathrm{ax}}/g_\pi =0`$ (solid line), (2) $`g_{\mathrm{ax}}/g_\pi =0.25`$ (dotted line), (3) $`g_{\mathrm{ax}}/g_\pi =0.5`$ (dashed line). The crosses indicate the points which correspond to $`m_\pi =140`$ MeV, and the diamonds indicate the points which correspond to $`m_\pi =280`$ MeV. 20. The numerical $`\mathrm{\Delta }_G`$ is plotted against the current quark mass $`m_0`$ for the three cases (1) $`g_{\mathrm{ax}}/g_\pi =0`$ (solid line), (2) $`g_{\mathrm{ax}}/g_\pi =0.25`$ (dashed line), (3) $`g_{\mathrm{ax}}/g_\pi =0.5`$ (dot-dot-dashed line). The analytic $`\mathrm{\Delta }_\mathrm{G}`$ is also plotted for these three cases (1) by dotted line, (2) by dot-dashed line, (3) by dot-dash-dashed line. The crosses indicate the points which correspond to $`m_\pi =140`$ MeV, and the diamonds indicate the points which correspond to $`m_\pi =280`$ MeV. 21. The violation of the PCAC relation $`\mathrm{\Delta }_\mathrm{P}`$ is plotted against the current quark mass $`m_0`$ for the three cases (1) $`g_{\mathrm{ax}}/g_\pi =0`$ (solid line), (2) $`g_{\mathrm{ax}}/g_\pi =0.25`$ (dotted line), (3) $`g_{\mathrm{ax}}/g_\pi =0.5`$ (dashed line). The crosses indicate the points which correspond to $`m_\pi =140`$ MeV, and the diamonds indicate the points which correspond to $`m_\pi =280`$ MeV.
warning/0004/cond-mat0004471.html
ar5iv
text
# Low energy collective modes, Ginzburg-Landau theory, and pseudogap behavior in superconductors with long-range pairing interactions ## I Introduction Superconductivity is a subject that has attracted prolonged interest in the condensed matter community. Much of our current understanding of this subject is based on the standard Bardeen-Cooper-Schrieffer (BCS) theory, which has enjoyed great success when applied to conventional superconductors. The discovery of the “high-$`T_c`$” cuprate superconductors opened a new chapter in the study of superconductivity. It is now well understood that just like in the BCS theory, electrons in cuprate superconductors form Cooper pairs. Moreover, it has been established that the wave functions of these Cooper pairs have predominantly $`d`$-wave symmetry. However, the origin of the force that leads to Cooper pairing in these systems is not yet understood and, experimentally, significant deviations from the BCS framework (which we take to include a Fermi liquid description of the normal state) have been found near and above the transition temperature, especially in the underdoped cuprates. Theoretically, there is a continuum of proposals for understanding these anomalies which range from conceptually modest (if practically far reaching) enhancements of the BCS framework (spin fluctuations, the crossover to pre-formed pairs, etc), to the influence of various hypothesized quantum critical points on to spin charge separation and non-Fermi liquid normal states. Finally there are scenarios based on self-organized dimensional reduction (stripes) which are very different in spirit. This paper, also motivated by the physics of the cuprates, is firmly in the conceptually modest camp. We examine a new crossover out of the BCS corner, this time to a superconductor with long range pairing interactions, i.e. in which the range of the pairing potential $`L`$ is much longer than the coherence length ($`L\xi `$) deduced from measurements of the (anisotropic) gap in the quasiparticle spectrum. We remind the reader that in the BCS theory, the phonon-mediated, retarded attraction between electrons is often modeled as a short-range instantaneous attractive potential. As a matter of fact, it is often modeled as a constant attractive scattering potential in momentum ($`𝐤`$) space, which corresponds to a $`\delta `$-function attraction in real space. This is an excellent approximation because in conventional superconductors the coherence length $`\xi `$ is of order $`1000\mathrm{\AA }`$, while the phonon-mediated attraction is a local (but retarded) on-site interaction with range of order the lattice spacing, and therefore the range is effectively zero compared to the size of the Cooper pairs. In such a short range model, the elementary excitations are quasiparticles with a ($`𝐤`$ independent) excitation gap $`\mathrm{\Delta }`$; the only collective mode below $`2\mathrm{\Delta }`$ is the linear Goldstone mode, whose energy is pushed up to the plasmon energy in real systems due to the existence of the long-range Coulomb interaction. The classic BCS limit also involves the assumption of weak coupling whereupon thermally excited quasiparticles control the largely mean-field superconducting transition temperature $`T_c`$, giving rise to a universal ratio $`2\mathrm{\Delta }/T_c3.5`$. Our interest in what happens when the range of the interaction is no longer the lattice spacing but instead crosses the (suitably defined, see below) coherence length arises from four considerations. First, $`\xi `$ is much shorter in the cuprates than in conventional superconductors; in fact $`\xi `$ is of order five lattice spacings in the a-b plane, and less than one lattice spacing along the c-axis. Therefore even if it turns out that the effective attractive interaction that gives rise to pairing is of the order of a lattice spacing, $`L`$ and $`\xi `$ will be comparable. Second, the physics of the pseudogap regime is most simply explained by invoking a suppression of $`T_c`$ by fluctuations that are small in the BCS limit. Quite generally the order parameter (or pair wavefunction) involves a relative piece and a center of mass piece, and the latter is what enters the standard Ginzburg-Landau theory. The simplest possibility is that the phase of the center of mass piece fluctuates strongly and this has been suggested in the context of the cuprates from two different standpoints, as the physics of strongly coupled pre-formed pairs or that of a low superfluid density coming from a doped Mott state. A more novel possibility is that the relative pair wavefunction has soft fluctuations, and as we demonstrate in this paper, these arise in the limit of a long range attraction. Third, it appears that the model we study here is relevant to some of the proposed mechanisms for cuprate superconductivity. The first example is the inter-layer pair hopping mechanism proposed by Anderson and co-workers. In this theory, pairing is induced by coherent pair hopping along the c-axis, provided that single electron hopping is frustrated by the non-Fermi liquid nature of the normal state. It was emphasized that high $`T_c`$ comes from the fact that the matrix element for pair hopping is diagonal in $`𝐤`$ space, although the symmetry of the order parameter is determined by the in-plane pairing potential. Mathematically, the pair hopping term is equivalent to an attractive potential off-diagonal in layer index, even though its physical origin is kinetic energy along the c-axis; in particular, being diagonal (or a $`\delta `$-function) in $`𝐤`$ space, it corresponds to a pairing potential that has an infinite range in real space. A second example of this is the spin-fluctuation mechanism proposed by Pines, Scalapino, and co-workers. In this theory, the range of the effective pairing interaction mediated by antiferromagnetic spin fluctuations is the spin-spin correlation length; as one reduces the doping level and approaches the antiferromagnetic instability from the overdoped side (where the system is more BCS-like), the spin-spin correlation length and therefore the range of the pairing interaction increases; it can become very large in the underdoped region, and diverges when approaching the antiferromagnetic instability. Finally, given the importance of the subject of superconductivity, we feel the situation we study here (which is opposite to the familiar short-range attraction limit) is interesting in its own right. Even if it proves not to be of particular relevance to superconductivity in cuprates, it may become relevant elsewhere. Our main results may be summarized as following. We find in addition to the usual quasiparticle excitations, the system supports collective modes whose energies are significantly lower than the quasiparticle gap $`\mathrm{\Delta }`$, when $`L\xi `$; in the limit that $`L\xi `$, the lowest energy collective mode gap becomes much lower than $`\mathrm{\Delta }`$, and the number of collective modes below $`2\mathrm{\Delta }`$ becomes large. At finite temperatures, due to the thermal fluctuations of these collective modes, the transition temperature $`T_c`$ becomes significantly lower than the mean field transition temperature, $`T_c^{MF}`$ (which is controlled by the thermally excited quasiparticles and does not know about the collective modes). For $`T_c<T<T_c^{MF}`$, the electrons are paired (in momentum space) but there is no long-range phase coherence; thus the system exhibits pseudogap behavior. Also the effective Ginzburg-Landau theory needs to be modified in this case; in addition to the center of mass degree of freedom of the superconducting order parameter (or Cooper pairs), which is the only degree of freedom taken into account in the original Ginsburg-Landau description, we also need to take into account the fluctuations of the relative or internal degrees of freedom of the order parameter. As a consequence spatial gradients couple to the internal pair structure in non-trivial ways, and although we have not done the computation yet, we suspect that the critical field $`H_{c2}`$ will be governed by a coherence length that is distinct from the one derived from the gap. In the main part of this paper, we will take the attitude that this is an interesting model system to study, and work out various properties of such a model, especially those that are different from the short-range limit of the BCS theory. Discussion of the possible relevance of our considerations to the cuprates will be reserved to the Summary section. The rest of the paper is organized as following. In Section II we revisit the Cooper problem, and keep the finite range of the attractive potential explicit. We note some problems involved in taking the limit of long-ranged potentials and specify the precise nature of the limit we consider in this paper. We find that in the limit $`L\xi `$, there are a large number of bound state solutions, whose binding energies are comparable to the ground state. These bound states correspond loosely to the collective modes in the many-body problem. In section III we study the BCS reduced Hamiltonian with a finite range attractive force, solving for both the mean-field ground state, and the zero momentum collective modes using the linearized equations of motion for the order parameter. In section IV we develop a Ginzburg-Landau effective theory from our model, using the functional integral method. We find that it is necessary to keep track of the fluctuations of both the internal and center of mass degrees of freedom of the superconducting order parameter; in particular, in the limit $`L\xi `$, $`T_c`$ is controlled by the fluctuations of the internal degrees of the order parameter, and is much lower than $`T_c^{MF}`$. In Section V we calculate the reduction of $`T_c`$ from $`T_c^{MF}`$ due to the thermal fluctuations of the collective modes, using the Ginzburg-Landau theory developed earlier. In Section VI we discuss the possible relevance of our work to the cuprate superconductors and mention some connections to other work in the literature. ## II The Cooper Problem We begin by studying the Cooper problem, in which the wave function of a Cooper pair takes the form: $$|\mathrm{\Psi }=\underset{0<ϵ_k<E_c}{}g(𝐤)c_𝐤^{}c_𝐤^{}|\mathrm{\Psi }_0,$$ (1) where $`|\mathrm{\Psi }_0`$ is the filled Fermi sea, $`ϵ_k=v_F(kk_F)`$ is the single electron energy measured from the Fermi level, and $`E_cE_F`$ (Fermi energy) is an energy cutoff. In phonon mediated attraction models, $`E_c`$ is usually taken to be the Debye energy. In this paper we treat $`E_c`$ as a parameter that can be varied in our effective model. The Hamiltonian (for both the Cooper problem considered here and the many-body problem in the BCS reduced Hamiltonian approximation considered in the following section) takes the form $$H=\underset{𝐤,\sigma }{}ϵ_kc_{𝐤\sigma }^{}c_{𝐤\sigma }\underset{\mathrm{𝐤𝐤}^{}}{}V_{|𝐤𝐤^{}|}c_𝐤^{}c_𝐤^{}c_𝐤^{}c_𝐤^{}.$$ (2) Here $$V_q=\frac{1}{A}d^2𝐫V(r)e^{i𝐪𝐫}$$ (3) is the Fourier transform of an attractive potential $`V(r)`$; $`A`$ is the area of the system. It is understood that the summation of the second term of Eq. (2) is restricted to states with $`|ϵ_k|,|ϵ_k^{}|<E_c`$. For simplicity, we assume $`V_q`$ and $`V(r)`$ are positive definite, so the ground state for the Cooper problem has $`s`$-wave symmetry, and the many-electron system is an $`s`$-wave superconductor. We will comment on the generalization to a d-wave superconductor in the concluding section. The Schroedinger equation that $`g(𝐤)`$ satisfies is $$Eg(𝐤)=2ϵ_kg(𝐤)\underset{𝐤^{}}{}V_{|𝐤𝐤^{}|}g(𝐤^{}).$$ (4) where $`E`$ is the eigenenergy. For a bound state, we must have $`E<0`$. For a general $`V_q`$, the integral equation (4) is difficult to solve. Cooper solved the short-range limit of this problem. In this limit, $`V_q=V`$ is a constant in momentum space, and therefore a $`\delta `$-function in real space. It was found that there exists only one bound state solution, with $`s`$-wave symmetry; for weak coupling ($`N(0)V1`$, $`N(0)`$ being the density of states at the Fermi energy), the energy of this state is $$E2E_ce^{2/N(0)V}.$$ (5) This well known dependence of the binding energy on $`V`$ is extremely singular. For finite but very short-range attractive potentials, there may be more bound state solutions (probably in other angular momentum channels). However, due to the singular dependence of the binding energy on the pairing potential, the binding energies of these states will be much smaller than the ground state, even if the effective pairing potential only changes slowly from one channel to another. Here we consider the opposite limit, in which $`V(r)`$ has a long but finite range, $`L`$. Although the explicit form of $`V(r)`$ is unimportant to our basic conclusions, for concreteness we assume it has a Gaussian form: $$V(r)=V_0e^{r^2/2L^2},$$ (6) and therefore $$V_q=\frac{2\pi L^2}{A}V_0e^{q^2L^2/2}.$$ (7) We are interested in following the evolution of the system as $`L`$ is increased and in the regime where $`L`$ is large, $`L\xi `$. One possibility is to consider potentials of the “Kac type” familiar from statistical mechanics, where the range is increased while keeping the integrated strength of the potential fixed, thereby achieving an unproblematic thermodynamic limit. A second possibility, perhaps more appropriate to attractive interactions generated by the system itself, is to keep the interaction of fixed magnitude ($`V_0`$) but to restrict its operation to an increasingly narrow shell around the Fermi surface. We will choose the second course here. An easy estimate shows that we need to pick the cutoff energy $$E_cL^D$$ (8) in order to keep the potential energy per particle finite. This also has the advantage that, by construction, is suppresses the tendency to phase separation that would otherwise be a complication with attractive long range interactions. There is however, a final caveat. For technical reasons, starting in the next paragraph, we will often employ a gradient expansion in momentum space. This is not quite consistent with the cutoff procedure but for the results of interest we will find that the precise value of the cutoff will enter weakly, leading us to believe that a better set of calculations will not alter our general scenario. Since $`V_{|𝐤𝐤^{}|}`$ goes to zero rapidly for $`|𝐤𝐤^{}|>1/L`$, if $`g(𝐤)`$ varies slowly on the scale of $`1/L`$, we may perform a gradient expansion for $`g(𝐤)`$ in $`𝐤`$-space, in the last term of Eq. (4): $`{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}g(𝐤^{})`$ $`=`$ $`{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}\left[g(𝐤)+_𝐤g(𝐤)(𝐤^{}𝐤)+{\displaystyle \frac{1}{2}}{\displaystyle \underset{\mu \nu }{}}{\displaystyle \frac{^2g(𝐤)}{k_\mu k_\nu }}(k_\mu ^{}k_\mu )(k_\nu ^{}k_\nu )+\mathrm{}\right]`$ (9) $``$ $`\left({\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}\right)g(𝐤)+{\displaystyle \frac{1}{4}}\left({\displaystyle \underset{𝐤^{}}{}}|𝐤𝐤^{}|^2V_{|𝐤𝐤^{}|}\right)_𝐤^2g(𝐤)`$ (10) $`=`$ $`V_0g(𝐤)+{\displaystyle \frac{V_0}{2L^2}}_𝐤^2g(𝐤),`$ (11) in which we have neglected higher gradient terms. Thus within the gradient expansion, Eq. (4) reduces to $$\frac{1}{2M}_𝐤^2g(𝐤)+(2ϵ_kV_0)g(𝐤)=Eg(𝐤),$$ (12) where $`M=L^2/V_0`$. This differential equation (12) is identical to the Schroedinger equation of a particle confined in an annulus (with some unknown, but calculable, boundary condition) with inner and outer radii $`k_F`$ and $`k_F+E_c/v_F`$ respectively, experiencing a “potential” $`U(k)=V_0+2ϵ_k=V_0+2v_F(kk_F)`$. It is interesting to note that in Eq. (12), the “kinetic energy” term comes from the two-body interaction in the original Hamiltonian, while the “potential” term actually is the original kinetic energy term. For large $`L`$, the “mass” $`M`$ is large, and we may use WKB approximation to analyze Eq. (12). The following conclusions follow straightforwardly: i) The ground state is in the $`s`$-wave channel, with energy $$E_0V_0+c(\frac{v_F\sqrt{V_0}}{L})^{2/3}V_0.$$ (13) Here $`c`$ is a constant of order 1; it is approximately $`(\frac{3\pi }{2\sqrt{2}})^{2/3}`$ for hard wall boundary conditions. We find the binding energy is essentially the depth of the attractive two-body potential $`V_0`$; this dependence is much less singular than the short-range limit defined above. Also the cutoff $`E_c`$ does not enter explicitly. At this point we define a “coherence length” $$\xi =2v_F/\pi |E_0|2v_F/\pi V_0,$$ (14) in analogy to the BCS coherence length at zero temperature: $`\xi _0=v_F/(\pi \mathrm{\Delta }(0))`$, where $`\mathrm{\Delta }(0)`$ is the quasiparticle gap at zero temperature. What we have in mind is that the binding energy in the Cooper problem will be the same as twice the quasiparticle gap in the BCS problem. This is not the case in short-range models; however, as we will see later, it is indeed true in the present case. We need to emphasize however, that $`\xi `$ is not the size of the Cooper pair wave function $`\mathrm{}`$ in the two-body problem studied in this section (this is the case even in short-range models); the latter can be estimated easily from Eq. (12). The size of the ground state wave function in momentum space for the Hamiltonian (12) is $$\mathrm{\Delta }k\left(\frac{1}{v_FM}\right)^{1/3}=\left(\frac{V_0}{v_FL^2}\right)^{1/3},$$ (15) thus $`\mathrm{}1/\mathrm{\Delta }k(v_FL^2/V_0)^{1/3}`$; it actually increases with $`L`$; however the sublinear dependence means for large $`L`$ we do have $`\mathrm{}L`$. This points to some ambiguities in what is meant by the coherence length in our problem and presumably spatial gradients may be governed by a different coherence length than the gap coherence length. We note that this possibility has been raised on completely different grounds in Ref. . In order for the gradient expansion performed above to be valid, we must have $`\mathrm{\Delta }k1/L`$, i.e., $`g(𝐤)`$ of the ground state varying slowly over the scale $`1/L`$, which is the range of $`V_q`$. This leads to the condition $$L2v_F/\pi V_0\xi ,$$ (16) as advertised earlier. It turns out that throughout this paper the gradient expansion in momentum space is the key technique that allows explicit calculations of various physical quantities to be made, and the above condition defines the range of its validity. For a given potential with fixed $`V_0`$ and $`L`$, one may also interpolate between the short-range and long-range limits by varying $`v_F`$. In the following we assume Eq. (16) is satisfied, and pay particular attention to the limit $`L\xi `$. ii) The number of bound states in the $`s`$-wave ($`l=0`$) channel is $`N_0\frac{\sqrt{2}L}{3\xi }`$. iii) The largest angular momentum channel that supports a bound state has angular momentum $`l_{max}\sqrt{2}k_FL`$. iv) The total number of bound states is $`N_{tot}=_lN_l\frac{\pi }{6}k_FL^2/\xi `$. v) The energy spacing between the ground state and the first excited state is $`\mathrm{\Delta }E\frac{V_0}{2k_F^2L^2}|E_0|`$, i.e., it is much lower than the binding energy of the ground state itself, in contrast to the short range case. The gradient expansion does not apply to all bound states, even if $`L\xi `$. One can show that it only applies to states with energy measured from the ground state (or bottom of the potential well) $`\mathrm{\Delta }EV_0\xi ^2/L^2`$. Nevertheless the estimate of number of bound states in various channels should be qualitatively correct. These large number of low-energy bound states correspond loosely to the low-energy collective modes in the many body problem studied in the following sections, and they lead to a significant reduction of $`T_c`$ from its mean-field value $`T_c^{MF}`$ (whose scale is set by the gap which is the same as the Cooper pair binding energy here, $`T_c^{MF}\mathrm{\Delta }V_0`$). This may be understood heuristically in the following way. When the temperature $`T`$ is reduced to $`T_c^{MF}`$, electrons start to form Cooper pairs. In the short-range BCS model, there is only one way to form a Cooper pair, thus all pairs condense into one state immediately and long-range coherence is established. In the model studied here however, there are many different ways to form Cooper pairs and their binding energies are comparable; thus the system needs to go to much lower temperature for all the Cooper pairs to condense into the ground state, hence $`T_cT_c^{MF}`$. For $`T_c<T<T_c^{MF}`$, the electrons are paired (and therefore gapped), but there is no long-range superconducting coherence, and the system exhibit pseudogap behavior. ## III Mean field theory and collective modes in the BCS reduced Hamiltonian In this section we study the BCS reduced Hamiltonian, Eq. (2). Instead of solving the two-body problem in the above section, here we study the many-body problem. The difference between Eq. (2) and the full many-body problem is that in Eq. (2), only pairs of electrons with opposite spin and total momentum zero scatter each other. (we return to the full problem, i.e. when the sum over the interacting momenta is constrained only by momentum conservation in the next section.) This leads to a very special property, that an unpaired electron never gets scattered, and thus the number of unpaired electron (0 or 1) for any $`𝐤`$ and spin orientation is a conserved quantity. The ground state for this Hamiltonian is in the subspace in which no such unpaired electron exists, namely the pair of single electron states $`𝐤`$ and $`𝐤`$ are either both occupied or both empty. In this subspace, one may map the problem onto a quantum pseudospin problem. We introduce a pseudospin-1/2 operator for each $`𝐤`$: $`S_𝐤^z`$ $`=`$ $`(c_𝐤^{}c_𝐤+c_𝐤^{}c_𝐤1)/2,`$ (17) $`S_𝐤^+`$ $`=`$ $`c_𝐤^{}c_𝐤^{},`$ (18) $`S_𝐤^{}`$ $`=`$ $`c_𝐤c_𝐤.`$ (19) In this mapping, the pseudospin is pointing up when a pair of single electron states are occupied, and down when they are empty; pair creation/annihilation operators map onto spin-flip operators. The Hamiltonian now takes the form (up to a constant): $$H=\underset{𝐤}{}2ϵ_𝐤S_𝐤^z\frac{1}{2}\underset{\mathrm{𝐤𝐤}^{}}{}V_{|𝐤𝐤^{}|}(S_𝐤^+S_𝐤^{}^{}+S_𝐤^{}^+S_𝐤^{})=\underset{𝐤}{}2ϵ_𝐤S_𝐤^z\underset{\mathrm{𝐤𝐤}^{}}{}V_{|𝐤𝐤^{}|}(S_𝐤^xS_𝐤^{}^x+S_𝐤^{}^yS_𝐤^y),$$ (20) i.e., the kinetic energy maps onto a Zeeman field that depends linearly on $`ϵ_𝐤`$ (and changes sign at the Fermi surface), which couples to the $`z`$ component of the pseudospins; the pairing interaction maps onto a ferromagnetic coupling among the $`xy`$ components of the pseudospins. In the mean field approximation, one replaces the ferromagnetic coupling among the $`xy`$ components of the spins by an average field: $$H_{MF}=\underset{𝐤}{}𝐁_𝐤^0𝐒_𝐤,$$ (21) where $`𝐁_𝐤^0`$ satisfies the self-consistency equation: $$𝐁_𝐤^0=2ϵ_𝐤\widehat{z}+2\underset{𝐤^{}}{}V_{|𝐤𝐤^{}|}𝐒__{}^{}{}_{𝐤}{}^{}^{}.$$ (22) Here $`𝐒__𝐤^{}`$ stands for the $`xy`$ components of $`𝐒_𝐤`$. In the ground state, $`𝐒_𝐤`$ points in the direction of $`𝐁_𝐤^0`$; the elementary excitations are single pseudospin flips, which corresponds to a pair of Bogliubov quasiparticles with opposite momenta and (real) spin. In the short-range limit $`L0`$, the range of $`V_{|𝐤𝐤^{}|}`$, $`1/L`$, diverges in $`𝐤`$ space. Thus all the pseudospins are equally coupled, no matter how far away they are in $`𝐤`$ space. In this limit, the mean field approximation becomes exact; the only collective mode is a zero mode corresponds to the global rotation of all the pseudospins along the $`z`$-axis, reflecting the broken XY symmetry of the ground state. All other excitations in this subspace may be described by pseudospin flips, or pairs of quasiparticles. The situation, however, becomes quite different, when $`L`$ becomes large. In this case $`V_{|𝐤𝐤^{}|}`$ becomes short-ranged in $`𝐤`$ space; one may actually divide the $`𝐤`$ space into blocks of size $`\frac{1}{L}\times \frac{1}{L}`$; within each block all the spins are strongly coupled and are effectively locked into a very big single spin; for different blocks, however, couplings are restricted to neighboring blocks. In such an $`XY`$ ferromagnet with “short range” couplings among the blocked spins, there are (pseudo)spin-wave like collective excitations in addition to single pseudospin flips, whose energies can become significantly lower than pseudospin flips for large $`L`$, as the number of blocks increases as $`L^2`$. We begin by analyzing the mean field equation, Eq. (22), at zero temperature. Without losing generality, we may assume $`𝐁_𝐤^0`$ is in the $`xz`$ plane: $$𝐁_𝐤^0=2ϵ_𝐤\widehat{z}+2\mathrm{\Delta }_𝐤\widehat{x},$$ (23) and the self-consistency equation for $`\mathrm{\Delta }_𝐤`$ is $$\mathrm{\Delta }_𝐤=\underset{𝐤^{}}{}V_{|𝐤𝐤^{}|}𝐒__𝐤^x=\frac{1}{2}\underset{𝐤^{}}{}\frac{V_{|𝐤𝐤^{}|}\mathrm{\Delta }_𝐤^{}}{\sqrt{ϵ_𝐤^{}^2+\mathrm{\Delta }_𝐤^{}^2}},$$ (24) which is the familiar BCS gap equation. In the limit $`L0`$, $`V_{|𝐤𝐤^{}|}=V`$ becomes independent of $`|𝐤𝐤^{}|`$, so does $`\mathrm{\Delta }_𝐤=\mathrm{\Delta }`$, and Eq. (24) reduces to $$1=\frac{1}{2}\underset{𝐤^{}}{}\frac{V}{\sqrt{ϵ_𝐤^{}^2+\mathrm{\Delta }^2}}.$$ (25) In the weak coupling limit, its solution is $`\mathrm{\Delta }=2E_ce^{1/N(0)V}`$. This gap is much larger than the binding energy in the Cooper problem, Eq. (5), due to the factor of two difference in the exponential. The situation becomes very different in the opposite limit, that $`L`$ becomes large. In this case $`V_{|𝐤𝐤^{}|}`$ becomes short-ranged in $`𝐤`$ space; thus on the right hand side of Eq. (24), we may replace $`\mathrm{\Delta }_𝐤^{}`$ and $`ϵ_𝐤^{}`$ by $`\mathrm{\Delta }_𝐤`$ and $`ϵ_𝐤`$ to first approximation, provided that they vary slowly on the scale of $`1/L`$. In this approximation one finds in the limit $`L\mathrm{}`$, $$\mathrm{\Delta }_𝐤=\sqrt{\frac{V_0^2}{4}ϵ_𝐤^2},$$ (26) for $`|ϵ_𝐤|<V_0/2`$, and $`\mathrm{\Delta }_𝐤=0`$ otherwise. For large but finite $`L`$, $`\mathrm{\Delta }_𝐤`$ is nonzero but exponentially small for $`|ϵ_𝐤|>V_0/2`$. Thus $`\mathrm{\Delta }_𝐤`$ is $`𝐤`$ dependent, and reaches its maximum for $`k=k_F`$, in which case $$\mathrm{\Delta }_{k_F}=\frac{V_0}{2}.$$ (27) We find the quasiparticle gap and the binding energy are indeed set by the same energy scale in the Cooper problem, as advertised earlier; in particular, the gap for creating a pair of quasiparticles on the Fermi surface, $`2\mathrm{\Delta }_{k_F}=V_0`$, is exactly the Cooper pair binding energy, in the limit $`L\mathrm{}`$. Following standard convention, we introduce the coherence length $$\xi =\frac{v_F}{\pi \mathrm{\Delta }_{k_F}}\frac{2v_F}{\pi V_0}.$$ (28) This definition indeed matches that of $`\xi `$ in the Cooper problem Eq. 14, in the limit that $`L`$ is large. And one can easily show that the large $`L`$ approximation is valid for $$L\xi .$$ (29) Within the mean field theory, the elementary excitations are the Bogliubov quasiparticles, with the standard spectrum $$E_k=\sqrt{ϵ_k^2+\mathrm{\Delta }_k^2}.$$ (30) At finite temperature, the self-consistent mean field equation becomes $$\mathrm{\Delta }_𝐤=\frac{1}{2}\underset{𝐤^{}}{}\frac{V_{|𝐤𝐤^{}|}\mathrm{\Delta }_𝐤^{}}{E_𝐤}\mathrm{tanh}\frac{\beta E_𝐤}{2},$$ (31) where $`\beta =\frac{1}{T}`$ is the inverse temperature (we set the Boltzmann constant $`k_B`$ to be 1). We can use the above equation to determine the mean-field transition temperature $`T_c^{MF}`$. In the $`L\mathrm{}`$ limit, it is particularly simple; in this case Eq. (31) reduces to $$1=\frac{V_0}{2E_𝐤}\mathrm{tanh}\frac{\beta E_𝐤}{2}.$$ (32) Using the facts that as $`TT_c^{MF}`$, $`\mathrm{\Delta }_𝐤0`$ and $`E_𝐤ϵ_𝐤`$, and focusing on $`|𝐤|k_F`$ where the pairing instability is the strongest, we find $$T_c^{MF}=\frac{V_0}{4}$$ (33) in this limit. Thus the ratio $`2\mathrm{\Delta }_{k_F}/T_c^{MF}=4`$, which is slightly bigger than the BCS ratio of $`3.5`$. There are a couple of new features of the mean-field solution in the large $`L`$ regime, that deserve some further discussion. (i) Unlike the case of short-range attraction originally considered by BCS, where the gap $`\mathrm{\Delta }`$ is a constant in momentum space at all temperatures (within the cutoff where the pairing interaction is nonzero), here $`\mathrm{\Delta }_𝐤`$ has strong momentum dependence, being maximum at the Fermi surface and decreasing as one moves away from it. At $`T=0`$, the momentum dependence Eq. (26) is such that the quasiparticle dispersion is essentially flat near the Fermi surface, for sufficiently large $`L`$. (ii) The temperature $`T`$ not only affects the overall scale of the gap, but also its momentum dependence. Taken at face value, one would conclude from Eq. (32) that $`T_c^{MF}`$ depends on $`ϵ_𝐤`$ in the following way: $$T_c^{MF}(ϵ_𝐤)=T_c^{MF}\frac{2ϵ_𝐤/V_0}{\mathrm{tanh}^1(2ϵ_𝐤/V_0)}T_c^{MF}.$$ (34) What this really means, of course, is that the gap $`\mathrm{\Delta }_𝐤`$ remains exponentially small for $`T_c^{MF}(ϵ_𝐤)T<T_c^{MF}`$. Thus in some sense $`T_c^{MF}`$ is a “local” property in the momentum space, increasing monotonically, and scaling roughly with the size of local gap $`\mathrm{\Delta }_𝐤(T=0)`$ unless $`\mathrm{\Delta }_𝐤(T=0)`$ is very small. In the following we go beyond mean field theory and study the collective excitations of the system at $`T=0`$. To do that, we study the equations of motion for $`𝐒_𝐤`$: $$\frac{d𝐒_𝐤}{dt}=i[𝐒_𝐤,H]=𝐒_𝐤\times 𝐁_𝐤,$$ (35) where $$𝐁_𝐤=2ϵ_𝐤\widehat{z}+2\underset{𝐤^{}}{}V_{|𝐤𝐤^{}|}𝐒__{}^{}{}_{𝐤}{}^{}^{}.$$ (36) Write $`𝐁_𝐤=𝐁_𝐤^0+\delta 𝐁_𝐤`$ and $`𝐒_𝐤=𝐒_𝐤^0+\delta 𝐒_𝐤`$, where $`𝐒_𝐤^0`$ is the expectation value of $`𝐒_𝐤`$ in the mean field ground state, and linearizing the equation of motion by neglecting terms proportional to $`\delta 𝐒_𝐤\times \delta 𝐁_𝐤`$, we obtain $$\frac{d\delta 𝐒_𝐤}{dt}=\delta 𝐒_𝐤\times 𝐁_𝐤^0+𝐒_𝐤^0\times \delta 𝐁_𝐤.$$ (37) $`\delta 𝐒_𝐤`$ should be perpendicular to $`𝐒_𝐤^0`$. If $`𝐒_𝐤^0`$ is in the $`xz`$ plane, we may assume $$\delta 𝐒_𝐤=\delta S_𝐤^y\widehat{y}+\delta S_𝐤^{}\widehat{e}_𝐤,$$ (38) where $`\widehat{e}_𝐤`$ is in the $`xz`$ plane, but perpendicular to $`𝐒_𝐤^0`$. Thus Eq. (37) reduces to $`{\displaystyle \frac{d\delta S_𝐤^y}{dt}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\delta B_𝐤^x\mathrm{cos}\theta _𝐤B_𝐤^0\delta S_𝐤^{},`$ (39) $`{\displaystyle \frac{d\delta S_𝐤^{}}{dt}}`$ $`=`$ $`B_𝐤^0\delta S_𝐤^y{\displaystyle \frac{1}{2}}\delta B_𝐤^y.`$ (40) Here $`\theta _𝐤`$ is the angle between $`𝐁_𝐤^0`$ and the $`\widehat{z}`$ direction, and $`\delta B_𝐤^x`$ $`=`$ $`2{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}\delta S_𝐤^{}^{}\mathrm{cos}\theta _𝐤^{},`$ (41) $`\delta B_𝐤^y`$ $`=`$ $`2{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}\delta S_𝐤^{}^y.`$ (42) Thus for a mode with frequency $`\omega `$ and $`\delta S_𝐤^y\varphi _𝐤`$, it must satisfy $`\omega ^2\varphi _𝐤`$ $`=`$ $`(B_𝐤^0)^2\varphi _𝐤+B_𝐤^0{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}\varphi _𝐤^{}+\mathrm{cos}\theta _𝐤{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}B_𝐤^{}^0\mathrm{cos}\theta _𝐤^{}\varphi _𝐤^{}`$ (43) $``$ $`\mathrm{cos}\theta _𝐤{\displaystyle \underset{𝐤^{}𝐤\mathrm{"}}{}}V_{|𝐤𝐤^{}|}V_{|𝐤^{}𝐤\mathrm{"}|}\mathrm{cos}\theta _𝐤^{}\varphi _{𝐤\mathrm{"}}.`$ (44) Performing a gradient expansion similar in spirit to the previous section, we obtain $`\omega ^2\varphi _𝐤`$ $`=`$ $`\stackrel{~}{U}_𝐤\varphi _𝐤{\displaystyle \frac{1}{4}}B_𝐤^0{\displaystyle \underset{𝐤^{}}{}}|𝐤𝐤^{}|^2V_{|𝐤𝐤^{}|}_𝐤^2\varphi _𝐤{\displaystyle \frac{\mathrm{cos}\theta _𝐤}{4}}({\displaystyle \underset{𝐤^{}}{}}|𝐤𝐤^{}|^2V_{|𝐤𝐤^{}|})_𝐤^2(\mathrm{cos}\theta _𝐤B_𝐤^0\varphi _𝐤)`$ (45) $`+`$ $`\mathrm{cos}\theta _𝐤[{\displaystyle \underset{𝐤^{}𝐤\mathrm{"}}{}}V_{|𝐤𝐤^{}|}V_{|𝐤^{}𝐤\mathrm{"}|}\mathrm{cos}\theta _𝐤^{}(𝐤\mathrm{"}𝐤)]_𝐤\varphi _𝐤`$ (46) $`+`$ $`{\displaystyle \frac{1}{2}}\mathrm{cos}\theta _𝐤[{\displaystyle \underset{𝐤^{}𝐤\mathrm{"}}{}}V_{|𝐤𝐤^{}|}V_{|𝐤^{}𝐤\mathrm{"}|}\mathrm{cos}\theta _𝐤^{}(k_\mu \mathrm{"}k_\mu )(k_\nu \mathrm{"}k_\nu )]_\mu _\nu \varphi _𝐤+\mathrm{},`$ (47) where the ellipses stand for terms that involve higher gradients which we neglect. Here $`\stackrel{~}{U}_𝐤`$ $`=`$ $`(B_𝐤^0)^2B_𝐤^0{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}\mathrm{cos}^2\theta _𝐤(B_𝐤^0{\displaystyle \underset{𝐤^{}}{}}V_{|𝐤𝐤^{}|}{\displaystyle \underset{𝐤^{}𝐤\mathrm{"}}{}}V_{|𝐤𝐤^{}|}V_{|𝐤^{}𝐤\mathrm{"}|}){\displaystyle \frac{V_0^2}{2L^2}}_𝐤^2\mathrm{cos}\theta _k`$ (48) $`=`$ $`(B_𝐤^0\mathrm{cos}^2\theta _𝐤V_0)(B_𝐤^0V_0){\displaystyle \frac{V_0^2}{2L^2}}_𝐤^2\mathrm{cos}\theta _k.`$ (49) In the limit that $`L`$ becomes very large, one finds $`\stackrel{~}{U}_𝐤=0`$ for $`|ϵ_𝐤|<V_0`$, and $`\stackrel{~}{U}_𝐤=(ϵ_𝐤V_0)^2`$ otherwise. For low energy modes, we expect $`\varphi _𝐤`$ to be centered near $`k=k_F`$. Thus in a somewhat crude approximation, we may neglect the gradient terms in Eq. (47) proportional to $`\mathrm{cos}\theta _𝐤`$, as $`\mathrm{cos}\theta _𝐤0`$ as $`kk_F`$; and assume $`\stackrel{~}{U}_𝐤`$ to become infinite for $`|ϵ_𝐤|>V_0`$. Within this approximation (which should give qualitatively correct results), we obtain the collective mode frequencies to be $$\omega _{mn}=\frac{V_0}{\sqrt{2}}\sqrt{\frac{m^2}{k_F^2L^2}+\frac{\pi ^4n^2}{(L/\xi )^2}},$$ (50) where $`m`$ and $`n`$ are “momentum”-like quantum numbers along and perpendicular to the Fermi surface respectively. So the mode frequency is linear in these “momenta”, as one would expect for the spin wave spectrum in an XY ferromagnet. It is clear that the energies of these collective modes become much lower than those of single pseudospin flips (corresponding to quasiparticle excitations), when $`L`$ is large. Since our discussion in this section is restricted to the BCS reduced Hamiltonian in which only pairs with zero total momentum interact with each other, these modes are exciton-like collective modes at zero momentum; they will acquire the usual dispersion of exciton modes at finite momentum. The term “exciton” indicates that the zero momentum excitation consists of a pair of quasiparticles bound into a pair state different from that of the condensate which is why the state lies in the gap between the ground state and the two quasiparticle continuum. The availability of such states (as we already saw in the previous section) is due to the long range of the interaction. ## IV Ginzburg-Landau theory and electromagnetic response In this section we use the functional integral formalism to derive the effective Ginzburg-Landau theory near their $`T_c^{MF}`$ for superconductors with finite range attractive interactions, and use it to derive their analog of the London equation. Let us consider the Hamiltonian $$\widehat{H}=\widehat{T}+\widehat{V}=\underset{k\sigma }{}(ϵ_k\mu )c_{𝐤\sigma }^{}c_{𝐤\sigma }𝑑𝐱𝑑𝐱^{}V(|𝐱𝐱^{}|)\mathrm{\Psi }_{}^{}(𝐱)\mathrm{\Psi }_{}^{}(𝐱^{})\mathrm{\Psi }_{}(𝐱^{})\mathrm{\Psi }_{}(𝐱).$$ (51) Here $`V(x)0`$ represents an attractive interaction. We will primarily be interested in singlet pairing in the ground state; so we neglect interactions between electrons with the same spin. While the interaction is written in a pairwise form, it is understood that a cutoff exists in momentum space so that only electrons that are close enough to the Fermi surface interact with each other, in the same manner as in previous sections. The difference here is that the interaction is no longer restricted to pairs of electrons with zero total momentum; thus the above Hamiltonian represents the full many-body problem, albeit with the parallel spin interaction ignored. One may also describe the system using an Euclidean action in terms of Grassman variables: $$S[\mathrm{\Psi },\overline{\mathrm{\Psi }}]=S_0[\mathrm{\Psi },\overline{\mathrm{\Psi }}]_0^\beta 𝑑\tau 𝑑𝐱𝑑𝐱^{}V(|𝐱𝐱^{}|)\overline{\mathrm{\Psi }}_{}(𝐱,\tau )\overline{\mathrm{\Psi }}_{}(𝐱^{},\tau )\mathrm{\Psi }_{}(𝐱^{},\tau )\mathrm{\Psi }_{}(𝐱,\tau ),$$ (52) where $`S_0`$ is the action for free electrons, and $`\tau `$ is the imaginary time. The partition function is $$Z=D\overline{\mathrm{\Psi }}D\mathrm{\Psi }e^{S[\mathrm{\Psi },\overline{\mathrm{\Psi }}]}.$$ (53) We now decouple the quartic term in $`S`$ by introducing a pair of Hubbard-Stratonovich fields $`\mathrm{\Delta }(𝐱,𝐱^{},\tau )`$ and $`\overline{\mathrm{\Delta }}(𝐱,𝐱^{},\tau )`$, which will become the superconducting order parameter in the sequel: $`S[\mathrm{\Psi },\overline{\mathrm{\Psi }},\mathrm{\Delta },\overline{\mathrm{\Delta }}]`$ $`=`$ $`S_0{\displaystyle _0^\beta }𝑑\tau {\displaystyle 𝑑𝐱𝑑𝐱^{}[\mathrm{\Delta }(𝐱,𝐱^{},\tau )\overline{\mathrm{\Psi }}_{}(𝐱,\tau )\overline{\mathrm{\Psi }}_{}(𝐱^{},\tau )+\overline{\mathrm{\Delta }}(𝐱,𝐱^{},\tau )\mathrm{\Psi }_{}(𝐱^{},\tau )\mathrm{\Psi }_{}(𝐱,\tau )]}`$ (54) $`+`$ $`{\displaystyle _0^\beta }𝑑\tau {\displaystyle 𝑑𝐱𝑑𝐱^{}\frac{|\mathrm{\Delta }(𝐱,𝐱^{},\tau )|^2}{V(|𝐱𝐱^{}|)}}.`$ (55) With this decoupling, the fermionic action becomes quadratic, and can be integrated out, after which we obtain an effective action in terms of the order parameter $`\mathrm{\Delta }(𝐱,𝐱^{},\tau )`$: $$S_e[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=_0^\beta 𝑑\tau 𝑑𝐱𝑑𝐱^{}\frac{|\mathrm{\Delta }(𝐱,𝐱^{},\tau )|^2}{V(|𝐱𝐱^{}|)}\mathrm{log}Z[\mathrm{\Delta },\overline{\mathrm{\Delta }}],$$ (56) where $$Z[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=D\overline{\mathrm{\Psi }}D\mathrm{\Psi }e^{S_0[\mathrm{\Psi },\overline{\mathrm{\Psi }}]+_0^\beta 𝑑\tau 𝑑𝐱𝑑𝐱^{}[\mathrm{\Delta }(𝐱,𝐱^{},\tau )\overline{\mathrm{\Psi }}_{}(𝐱,\tau )\overline{\mathrm{\Psi }}_{}(𝐱^{},\tau )+\overline{\mathrm{\Delta }}(𝐱,𝐱^{},\tau )\mathrm{\Psi }_{}(𝐱^{},\tau )\mathrm{\Psi }_{}(𝐱,\tau )]}.$$ (57) The mean-field solution corresponds to the saddle point of $`S_e[\mathrm{\Delta },\overline{\mathrm{\Delta }}]`$: $$\frac{\delta S_e[\mathrm{\Delta },\overline{\mathrm{\Delta }}]}{\delta \overline{\mathrm{\Delta }}(𝐱,𝐱^{})}|{}_{\mathrm{\Delta }=\mathrm{\Delta }_s}{}^{}=\frac{\mathrm{\Delta }_s(𝐱,𝐱^{})}{V(𝐱𝐱^{})}\frac{\delta \mathrm{log}Z[\mathrm{\Delta },\overline{\mathrm{\Delta }}]}{\delta \overline{\mathrm{\Delta }}(𝐱,𝐱^{})}|_{\mathrm{\Delta }=\mathrm{\Delta }_s}=\frac{\mathrm{\Delta }_s(𝐱,𝐱^{})}{V(𝐱𝐱^{})}\mathrm{\Psi }_{}(𝐱^{})\mathrm{\Psi }_{}(𝐱)_{\mathrm{\Delta }_s}=0,$$ (58) where $`_{\mathrm{\Delta }_s}`$ stands for quantum and thermal averaging in the presence of the pairing field $`\mathrm{\Delta }_s(𝐱,𝐱^{})`$. Here we have assumed a static saddle point so that $`\mathrm{\Delta }_s`$ has no $`\tau `$ dependence. Further assuming that at the saddle point, $`\mathrm{\Delta }_s(𝐱,𝐱^{})=\mathrm{\Delta }_s(𝐱𝐱^{})`$ is translationally invariant, it is easy to show that Eq. (58) is equivalent to Eq. (24), upon the identification $$\mathrm{\Delta }_𝐤=\frac{1}{A}𝑑𝐱e^{i𝐤𝐱}\mathrm{\Delta }_s(𝐱).$$ (59) The functional integral formalism can be used to derive the effective Ginzburg-Landau free energy, in the vicinity of $`T_c^{MF}`$. This has been done for the short-range attractive interactions. Here we use it to derive the appropriate Ginzburg-Landau free energy for finite range attractive interactions. Our starting point is the effective action, Eq. (56). Near $`T_c^{MF}`$, we may make two simplifications: i) We may neglect the $`\tau `$ dependence of $`\mathrm{\Delta }`$ as we expect the thermal fluctuations to dominate the quantum fluctuations; ii) We may expand $`S_e`$ in powers of $`\mathrm{\Delta }`$. The quadratic terms take the form $$S_e^{(2)}[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=\beta 𝑑𝐱𝑑𝐲\frac{|\mathrm{\Delta }(𝐱,𝐲)|^2}{V(|𝐱𝐲|)}𝑑𝐱_1𝑑𝐲_1𝑑𝐱_2𝑑𝐲_2Q(𝐱_1,𝐲_1;𝐱_2,𝐲_2)\mathrm{\Delta }(𝐱_1,𝐲_1)\overline{\mathrm{\Delta }}(𝐱_2,𝐲_2),$$ (60) where $`Q`$ $`(𝐱_1,𝐲_1;𝐱_2,𝐲_2)={\displaystyle \frac{\delta ^2\mathrm{log}Z}{\delta \mathrm{\Delta }(𝐱_1,𝐲_1)\delta \overline{\mathrm{\Delta }}(𝐱_2,𝐲_2)}}|_{\mathrm{\Delta }=0}`$ (61) $`=`$ $`{\displaystyle _0^\beta }𝑑\tau _1{\displaystyle _0^\beta }𝑑\tau _2\mathrm{\Psi }_{}(𝐲_2,\tau _2)\mathrm{\Psi }_{}(𝐱_2,\tau _2)\overline{\mathrm{\Psi }}_{}(𝐱_1,\tau _1)\overline{\mathrm{\Psi }}_{}(𝐲_1,\tau _1)_c`$ (62) $`=`$ $`{\displaystyle _0^\beta }𝑑\tau _1{\displaystyle _0^\beta }𝑑\tau _2G_0(𝐲_2𝐲_1;\tau _2\tau _1)G_0(𝐱_2𝐱_1;\tau _2\tau _1)`$ (63) $`=`$ $`{\displaystyle \underset{i\omega _n}{}}G_0(𝐲_2𝐲_1;i\omega _n)G_0(𝐱_2𝐱_1;i\omega _n).`$ (64) Here $`G_0`$ is the normal state (non-interacting) single electron Green’s function, $`\omega _n`$’s are fermion Masubara frequencies, and $`_c`$ stands for connected contractions in the average. We now introduce “center of mass” and “relative” coordinates for the order parameter $`\mathrm{\Delta }(𝐱,𝐲)`$: $$𝐑=(𝐱+𝐲)/2,𝐫=𝐲𝐱,$$ (65) and Fourier transform with respect to the relative coordinate $`𝐫`$: $$\mathrm{\Delta }(𝐑,𝐤)=𝑑𝐫e^{i𝐤𝐫}\mathrm{\Delta }(𝐑𝐫/2,𝐑+𝐫/2).$$ (66) In a uniform ($`𝐑`$ independent) configuration, $`\mathrm{\Delta }(𝐑,𝐤)`$ becomes $`\mathrm{\Delta }_𝐤`$. In terms of $`\mathrm{\Delta }(𝐑,𝐤)`$, the first term of Eq. (60) becomes $`\beta `$ $`{\displaystyle 𝑑𝐱𝑑𝐲\frac{|\mathrm{\Delta }(𝐱,𝐲)|^2}{V(|𝐱𝐲|)}}=\beta {\displaystyle 𝑑𝐑𝑑𝐫\frac{|\mathrm{\Delta }(𝐑\frac{𝐫}{2},𝐑+\frac{𝐫}{2})|^2}{V(r)}}`$ (67) $`=`$ $`{\displaystyle \frac{\beta }{A^2}}{\displaystyle 𝑑𝐑\underset{𝐤_1𝐤_2}{}\mathrm{\Delta }(𝐑,𝐤_1)\overline{\mathrm{\Delta }}(𝐑,𝐤_2)F(|𝐤_1𝐤_2|)},`$ (68) where $$F(|𝐤_1𝐤_2|)=𝑑𝐫e^{i(𝐤_1𝐤_2)𝐫}/V(r).$$ (69) This expression is problematic as it stands, for $`1/V(r)`$ does not, in a strict sense, possess a Fourier transform in a truly infinite system. However, it can be defined in a large but finite-size system with periodic boundary conditions; and as we will see in a few lines, this is only an intermediate expression and can be regulated as such and the final answer will be sensible even as the regulator is removed. It is easy to see that such a regulated $`F(k)`$ must be peaked near $`k=0`$, and therefore short-ranged in $`k`$ space. Thus we may perform a gradient expansion for $`\mathrm{\Delta }(𝐑,𝐤)`$ in $`k`$ space, as in previous sections. Introducing $`𝐤=(𝐤_1+𝐤_2)/2`$, $`𝐤^{}=𝐤_2𝐤_1`$, we obtain $`{\displaystyle \frac{\beta }{A^2}}`$ $`{\displaystyle 𝑑𝐑\underset{𝐤_1𝐤_2}{}\mathrm{\Delta }(𝐑,𝐤_1)\overline{\mathrm{\Delta }}(𝐑,𝐤_2)F(|𝐤_1𝐤_2|)}`$ (70) $`=`$ $`{\displaystyle \frac{\beta }{A^2}}{\displaystyle 𝑑𝐑\underset{\mathrm{𝐤𝐤}^{}}{}(|\mathrm{\Delta }(𝐑,𝐤)|^2\frac{|𝐤^{}|^2}{4}|_𝐤\mathrm{\Delta }(𝐑,𝐤)|^2+\mathrm{})F(k^{})}`$ (71) $`=`$ $`{\displaystyle \frac{\beta }{A}}{\displaystyle \underset{𝐤}{}}{\displaystyle 𝑑𝐑[B_1|\mathrm{\Delta }(𝐑,𝐤)|^2+B_2|_𝐤\mathrm{\Delta }(𝐑,𝐤)|^2+\mathrm{}]},`$ (72) where $`B_1`$ $`=`$ $`{\displaystyle \frac{1}{A}}{\displaystyle \underset{𝐤}{}}F(k)={\displaystyle \frac{1}{V(0)}},`$ (73) $`B_2`$ $`=`$ $`{\displaystyle \frac{1}{A}}{\displaystyle \underset{𝐤}{}}F(k){\displaystyle \frac{k^2}{4}}={\displaystyle \frac{1}{4}}^2({\displaystyle \frac{1}{V(r)}})|_{r=0}.`$ (74) For $`V(r)=V_0e^{r^2/2L^2}`$, we have $`B_1`$ $`=`$ $`{\displaystyle \frac{1}{V_0}},`$ (75) $`B_2`$ $`=`$ $`{\displaystyle \frac{1}{2V_0L^2}}.`$ (76) Clearly, the coefficient $`B_2`$ controls the fluctuations of $`\mathrm{\Delta }`$ in the $`𝐤`$ space, which describes the internal degrees of freedom for the pairing amplitude; the larger the interaction range $`L`$ is, the softer such fluctuations are. On the other hand, in the short range limit $`L0`$, such fluctuations get completely suppressed, and the order parameter $`\mathrm{\Delta }`$ depends only of the center of mass coordinate $`𝐑`$, and we recover the standard Ginzburg-Landau theory in terms of $`\mathrm{\Delta }(𝐑)`$ with no $`𝐤`$ dependence. It is clear however, we need to keep full $`𝐤`$ dependence of $`\mathrm{\Delta }`$ in the present case. In the first term of Eq. (60), there is no mechanism to control the spatial ($`𝐑`$) fluctuations of $`\mathrm{\Delta }`$. Such fluctuations are controlled by the second term, which we now turn to. The second term in Eq. (60) takes the form $``$ $`{\displaystyle 𝑑𝐱_1𝑑𝐲_1𝑑𝐱_2𝑑𝐲_2Q(𝐱_1,𝐲_1;𝐱_2,𝐲_2)\mathrm{\Delta }(𝐱_1,𝐲_1)\overline{\mathrm{\Delta }}(𝐱_2,𝐲_2)}`$ (77) $`=`$ $`{\displaystyle \frac{1}{A}}{\displaystyle \underset{𝐤}{}}{\displaystyle 𝑑𝐑_1𝑑𝐑_2\mathrm{\Delta }(𝐑_1,𝐤)\overline{\mathrm{\Delta }}(𝐑_2,𝐤)P(𝐑_2𝐑_1,𝐤)},`$ (78) where $$P(𝐑_2𝐑_1,𝐤)=\frac{1}{A}\underset{𝐤_1,i\omega _n}{}G_0(𝐤_1,i\omega _n)G_0(2𝐤+𝐤_1,i\omega _n)e^{i(2𝐤+2𝐤_1)(𝐑_2𝐑_1)},$$ (79) and $$G_0(𝐤,i\omega _n)=\frac{1}{i\omega _nϵ_𝐤}.$$ (80) In this term there is coupling between $`\mathrm{\Delta }`$’s with different $`𝐑`$’s, but no coupling between $`\mathrm{\Delta }`$’s with different $`𝐤`$’s. Thus the first and second quadratic terms in the action control the internal (k) and spatial ($`𝐑`$) fluctuations of the order parameter $`\mathrm{\Delta }`$ respectively. The fluctuations of the overall magnitude of the order parameter are controlled by higher order terms in the effective action. As usual we may stop at the quartic term near $`T_c^{MF}`$, and neglect terms that involve the spatial gradient of the order parameter at this order. For the present problem, we obtain (in a way similar to Ref. ) $$S_e^{(4)}=\frac{1}{8A}\underset{𝐤}{}𝑑𝐑|\mathrm{\Delta }(𝐑,𝐤)|^4(\frac{\mathrm{tanh}(\beta ϵ_𝐤/2)}{ϵ_𝐤^3}\frac{\beta }{2ϵ_𝐤^2\mathrm{cosh}^2(\beta ϵ_𝐤/2)}).$$ (81) Now we perform a spatial gradient expansion in $`𝐑`$, for the second quadratic term, Eq. (78). To do that, we first go to momentum space, and define $$\mathrm{\Delta }(𝐐,𝐤)=𝑑𝐑e^{i𝐐𝐑}\mathrm{\Delta }(𝐑,𝐤).$$ (82) In terms of $`\mathrm{\Delta }(𝐐,𝐤)`$, the second quadratic term reads $$\frac{1}{A^2}\underset{𝐐,𝐤}{}D(𝐐,𝐤)|\mathrm{\Delta }(𝐐,𝐤)|^2,$$ (83) where $`D(𝐐,𝐤)`$ $`=`$ $`{\displaystyle \underset{i\omega _n}{}}G_0(𝐐/2𝐤,i\omega _n)G_0(𝐐/2+𝐤,i\omega _n)`$ (84) $`=`$ $`{\displaystyle \frac{\beta }{2}}{\displaystyle \frac{\mathrm{tanh}(\beta ϵ_{𝐐/2𝐤}/2)+\mathrm{tanh}(\beta ϵ_{𝐐/2+𝐤}/2)}{ϵ_{𝐐/2𝐤}+ϵ_{𝐐/2+𝐤}}}.`$ (85) Expanding for small $`Q`$, we obtain $$D(𝐐,𝐤)D(0,𝐤)\frac{\beta ^3v_F^2\mathrm{sinh}(\beta ϵ_𝐤/2)}{32ϵ_𝐤\mathrm{cosh}^3(\beta ϵ_𝐤/2)}(𝐐\widehat{𝐤})^2.$$ (86) Fourier transforming back to real ($`𝐑`$) space, we obtain $$\frac{1}{A}\underset{𝐤}{}D(0,𝐤)𝑑𝐑|\mathrm{\Delta }(𝐑,𝐤)|^2+\frac{1}{A}\underset{𝐤}{}\frac{\beta ^3v_F^2\mathrm{sinh}(\beta ϵ_𝐤/2)}{32ϵ_𝐤\mathrm{cosh}^3(\beta ϵ_𝐤/2)}𝑑𝐑|\widehat{𝐤}\mathrm{\Delta }(𝐑,𝐤)|^2.$$ (87) Thus putting all these terms together, and keeping leading gradient terms only, we obtain $`S_e[\mathrm{\Delta },\overline{\mathrm{\Delta }}]`$ $`=`$ $`{\displaystyle \frac{\beta }{A}}{\displaystyle \underset{𝐤}{}}{\displaystyle }d𝐑\{[B_1C(\beta ,𝐤)]|\mathrm{\Delta }(𝐑,𝐤)|^2+B_2|_𝐤\mathrm{\Delta }(𝐑,𝐤)|^2`$ (88) $`+`$ $`{\displaystyle \frac{1}{2m_𝐤}}|\widehat{𝐤}\mathrm{\Delta }(𝐑,𝐤)|^2+U(k)|\mathrm{\Delta }(𝐑,𝐤)|^4\},`$ (89) where $`B_1`$ $`=`$ $`1/V_0,`$ (90) $`B_2`$ $`=`$ $`1/(2V_0L^2),`$ (91) $`C(\beta ,k)`$ $`=`$ $`\mathrm{tanh}{\displaystyle \frac{\beta ϵ_k}{2}}/(2ϵ_k),`$ (92) $`{\displaystyle \frac{1}{2m_𝐤}}`$ $`=`$ $`{\displaystyle \frac{\beta ^2v_F^2\mathrm{sinh}\frac{\beta ϵ_k}{2}}{32ϵ_k\mathrm{cosh}^3\frac{\beta ϵ_k}{2}}},`$ (93) $`U(k)`$ $`=`$ $`{\displaystyle \frac{\mathrm{tanh}\frac{\beta ϵ_k}{2}}{8ϵ_k^3}}{\displaystyle \frac{\beta }{16ϵ_k^2\mathrm{cosh}^2\frac{\beta ϵ_k}{2}}}.`$ (94) Eq. (89) is one of the central results of this paper; it is the basis for the calculation of reduction of $`T_c`$ from $`T_c^{MF}`$ due to thermal fluctuations of collective modes in the next section. Thus far in our discussion we have assumed a uniform space with no background electromagnetic field. In the presence of a weak and slowly varying background magnetic field, using standard arguments one finds that the only modification one needs to make in (89) is to replace $``$ by $`+2ie𝐀(𝐑)/c`$. The supercurrent is $$𝐣(𝐑)=c\frac{\delta S_e}{\delta 𝐀(𝐑)}=\frac{ie\beta }{A}\underset{𝐤}{}\frac{𝐤}{m_k}\{\overline{\mathrm{\Delta }}[𝐤(+\frac{2ie𝐀}{c})\mathrm{\Delta }]c.c.\}.$$ (95) ¿From this we may obtain the equivalent of London’s equation: $$𝐣(𝐑)=\frac{4\beta e^2}{Ac}\underset{𝐤}{}|\mathrm{\Delta }_𝐤|^2\frac{𝐤}{m_k}[𝐤𝐀(𝐑)]=\{\frac{4e^2}{Acd}\underset{𝐤}{}\frac{|\mathrm{\Delta }_𝐤|^2k^2}{m_k}\}𝐀(𝐑),$$ (96) from which the expression for penetration depth follows: $$\frac{1}{\lambda ^2}=\frac{8\beta e^2}{Adc^2}\underset{𝐤}{}\frac{|\mathrm{\Delta }_𝐤|^2k^2}{m_k}.$$ (97) Here $`d`$ is the spacing between layers. These are, of course, mean-field results; fluctuations have been left out at this level. And they apply only near $`T_c^{MF}`$, even at the mean-field level. ## V reduction of $`T_c`$ due to thermal fluctuations of collective modes and pseudogap behavior We study in this section the reduction of of $`T_c`$ from $`T_c^{MF}`$ due to the fluctuations of the collective modes, using the Ginzburg-Landau free energy derived earlier. As a warm-up, as well as for the purpose of later comparison, we study the same effect in a weak coupling BCS superconductor with short range ($`\delta `$-function) interaction, whose Ginzburg-Landau free energy is the familiar $`O(N)`$ $`\varphi ^4`$ theory with $`N=2`$: $$S_e=\beta 𝑑𝐫\left(\frac{1}{4m_e}|\psi |^2+a|\psi |^2+b|\psi |^4\right),$$ (98) where in 3D we have $`\psi (𝐫)`$ $`=`$ $`\left({\displaystyle \frac{7\zeta (3)m_ek_Fϵ_F}{12\pi ^4(T_c^{MF})^2}}\right)^{1/2}\mathrm{\Delta }(𝐫),`$ (99) $`a`$ $`=`$ $`{\displaystyle \frac{6\pi ^2T_c^{MF}(TT_c^{MF})}{7\zeta (3)ϵ_F}},`$ (100) $`b`$ $`=`$ $`{\displaystyle \frac{9\pi ^4(T_c^{MF})^2}{14\zeta (3)m_ek_Fϵ_F^2}},`$ (101) and $`m_e`$ is the effective mass of the electron. In the mean field theory, one neglects the quartic term in (101), and $`T_c`$ is determined by the point where $`a`$ turns negative, which is nothing but $`T_c^{MF}`$. The fluctuation effects of the quartic term may be studied using the self-consistent field approximation that is exact in the limit $`N\mathrm{}`$. Within this approximation, one replaces $`|\psi |^4`$ by $`4|\psi |^2|\psi |^2`$, and requires that the following self-consistency equation is satisfied: $$|\psi |^2=\frac{1}{V}\underset{𝐤}{}\psi _𝐤\psi _𝐤=\frac{1}{(2\pi )^3}d^3𝐤\frac{k_BT}{\frac{k^2}{4m_e}+a+4b|\psi |^2}.$$ (102) At $`T=T_c`$, we ought to have $`a+4b|\psi |^2=0`$. Thus the equation that determines $`T_c`$ becomes $$a+4b\frac{4m_ek_BT_c}{(2\pi )^3}_{k<\mathrm{\Lambda }_{uv}}\frac{d^3𝐤}{k^2}=0,$$ (103) where $`\mathrm{\Lambda }_{uv}\mathrm{}\omega _D/v_F`$ is an appropriate high momentum (ultraviolet) cutoff for $`k`$. From this we obtain $$\frac{T_c^{MF}T_c}{T_c^{MF}}\frac{k_BT_c\mathrm{}\omega _D}{ϵ_F^2}1$$ (104) for weak coupling superconductors, whence it is clear that the mean field value for $`T_c`$ is extremely accurate. In the following we demonstrate that the situation is very different, and $`T_c`$ becomes significantly lower than $`T_c^{MF}`$ when $`L\xi `$, due to thermal fluctuations of the low energy collective modes discussed in previous sections. We use the Ginzburg-Landau theory appropriate for this situation, Eq. (89), and use the same self-consistent field approximation as above by replacing $`|\mathrm{\Delta }(𝐑,𝐤)|^4`$ with $`4|\mathrm{\Delta }(𝐑,𝐤)|^2|\mathrm{\Delta }(𝐑,𝐤)|^2`$, after which the reduced action takes a quadratic form: $$S_R[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=\frac{\beta }{A}\underset{𝐤}{}𝑑𝐑\left\{[B_1C(\beta ,𝐤)+\stackrel{~}{U}(k)]|\mathrm{\Delta }(𝐑,𝐤)|^2+B_2|_𝐤\mathrm{\Delta }(𝐑,𝐤)|^2+\frac{1}{2m_𝐤}|\widehat{𝐤}\mathrm{\Delta }(𝐑,𝐤)|^2\right\},$$ (105) where $$\stackrel{~}{U}(k)=4U(k)|\mathrm{\Delta }(𝐑,𝐤)|^2.$$ (106) We would like to formally diagonalise this quadratic form. To do that, we take advantage of the translation invariance of the action, and express the reduced action in terms of the Fourier transform of $`\mathrm{\Delta }(𝐑,𝐤)`$, $`\mathrm{\Delta }(𝐐,𝐤)`$: $$S_R[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=\frac{\beta }{A^2}\underset{𝐤,𝐐}{}\left\{\left[B_1C(\beta ,𝐤)+\stackrel{~}{U}(k)+\frac{(\widehat{𝐤}𝐐)^2}{2m_k}\right]|\mathrm{\Delta }(𝐐,𝐤)|^2+B_2|_𝐤\mathrm{\Delta }(𝐐,𝐤)|^2\right\}.$$ (107) Note that modes with different $`𝐐`$’s decouple. To proceed further, we introduce eigen modes $`\psi _{mn}(𝐐,𝐤)`$ which satisfy $$\left[B_1C(\beta ,𝐤)+\stackrel{~}{U}(k)+\frac{(\widehat{𝐤}𝐐)^2}{2m_k}B_2_𝐤^2\right]\psi _{mn}(𝐐,𝐤)=E_{mn}(Q)\psi _{mn}(𝐐,𝐤),$$ (108) and the normalization condition: $$\frac{1}{A}\underset{𝐤}{}|\psi _{mn}(𝐐,𝐤)|^2=1.$$ (109) Here $`m`$ and $`n`$ are quantum numbers to be specified later. Expanding $`\mathrm{\Delta }(𝐐,𝐤)`$ in terms of $`\psi _{mn}(𝐐,𝐤)`$: $$\mathrm{\Delta }(𝐐,𝐤)=\underset{mn}{}a_{mn}(𝐐)\psi _{mn}(𝐐,𝐤),$$ (110) we bring the reduced action to diagonal form: $$S_R[\mathrm{\Delta },\overline{\mathrm{\Delta }}]=\frac{\beta }{A}\underset{𝐐mn}{}E_{mn}(Q)|a_{mn}(𝐐)|^2.$$ (111) The self-consistent equation now becomes $$\stackrel{~}{U}(k)=4U(k)|\mathrm{\Delta }(𝐑,𝐤)|^2=\frac{4U(k)}{A^2}\underset{𝐐mn}{}|\psi _{mn}(𝐐,𝐤)|^2|a_{mn}(𝐐)|^2=\frac{4U(k)}{A\beta }\underset{𝐐mn}{}\frac{|\psi _{mn}(𝐐,𝐤)|^2}{E_{mn}(Q)}.$$ (112) At $`T=T_c`$, we have the lowest eigenvalue $$E_{00}(Q=0)=0;$$ (113) This poses a self-consistent condition that determines $`T_c`$. The Schroedinger-like equation (108) is quite difficult to solve in general, one of the reason being the isotropy in $`𝐤`$ space is lost due to the presence of the term proportional to $`(\widehat{𝐤}𝐐)^2`$. In order to proceed, we need to make a number of further approximations and assumptions. Firstly, we average $`(\widehat{𝐤}𝐐)^2`$ along different directions of $`𝐤`$ and replace it by $`Q^2/2`$, so that isotropy in $`𝐤`$ space is restored and all modes may be labeled by an “angular momentum” quantum number $`m`$. We also notice that $`C(\beta ,k)`$ has its minimum at $`k=k_F`$, and goes to zero rapidly for $`ϵ_k>1/\beta `$. Hence we may approximate it by a “square” well, with depth $`C(\beta ,k_F)=\frac{\beta }{4}`$ and width $`\frac{2}{v_F\beta }`$. $`U(k)`$ (and therefore $`\stackrel{~}{U}(k)`$) has a similar structure, and we can also approximate it by a step-like structure, with height $`U_{k_F}=\frac{\beta ^3}{96}`$ and the same width $`\frac{2}{v_F\beta }`$. Anticipating that the “low-energy” ($`E`$) modes will be localized in the step well, we may approximate $`\frac{1}{2m_k}`$ by a constant $`\frac{1}{2m_{k_F}}`$. With these simplifications, Eq. (108) reduces to that of the Schroedinger equation of a particle confined to a step potential well. Imposing periodic boundary condition at the ends of the well, we obtain $$\psi _{mn}(𝐐,𝐤)=\sqrt{\frac{\pi v_F\beta }{k_F}}e^{i[2\pi m\theta _𝐤+n\pi (kk_F)/v_F\beta ]},$$ (114) and at $`T=T_c`$ (using (113)) $$E_{mn}(Q)=\frac{\beta ^3v_F^2Q^2}{64}+B_2(\frac{m^2}{k_F^2}+\pi ^2n^2v_F^2\beta ^2).$$ (115) Combining Eqs. (112) and (115), we obtain the equation that determines $`T_c`$: $$B_1\frac{\beta _c}{4}+\frac{2}{3\pi k_Fv_F}\underset{mn}{}\frac{d𝐐}{Q^2+\frac{64B_2}{\beta _cv_F^2k_F^2}(m^2+\pi ^2n^2\beta _c^2v_F^2k_F^2)}=0.$$ (116) We notice that for the case of $`m=0`$ and $`n=0`$, the integral is logarithmically divergent at both infrared and ultraviolet. As discussed earlier, there is always an ultraviolet cutoff $`\mathrm{\Lambda }_{uv}E_c/v_F`$. The infrared divergence is a signature of the fact that 2D is the lower dimensionality for ordering of this model (we do not go into details of the Kosterlitz-Thouless picture for a true phase transition for $`N=2`$ here). An infrared cutoff $`\mathrm{\Lambda }_{ir}`$ is provided by invoking the quasi-2D nature of all real systems, which eventually crossover to 3D at sufficiently long length scales. Assuming $`\mathrm{\Lambda }_{ir}`$ is sufficiently large compared to the energies (measured in proper units) of the modes whose fluctuations contribute significantly to the reduction of $`T_c`$, we obtain $$\beta _c\frac{4M}{3v_Fk_F}\mathrm{log}\frac{\mathrm{\Lambda }_{uv}}{\mathrm{\Lambda }_{ir}},$$ (117) where $`M`$ is the number of modes contributing in the sum of $`m`$ and $`n`$. To determine $`M`$, we note that the self-consistent potential (for $`Q=0`$; finite $`Q`$ only adds a constant to it) is zero inside the well, and $`B_1`$ outside it. Thus in order for the previous approximations for mode solutions and “energies” to be valid, we need to have $`E_{mn}(Q=0)<B_1`$. Summing up the number of these modes, we obtain $$Mk_F\sqrt{B1/B_2}=k_FL.$$ (118) We thus find $$T_c=k_B/\beta _c\frac{v_F}{L\mathrm{log}\frac{\mathrm{\Lambda }_{uv}}{\mathrm{\Lambda }_{ir}}}\frac{T_c^{MF}}{(L/\xi )\mathrm{log}\frac{\mathrm{\Lambda }_{uv}}{\mathrm{\Lambda }_{ir}}}.$$ (119) It clearly goes to zero as $`L\mathrm{}`$. We would like to emphasize that here we have made a number of crude approximations in the calculation of $`T_c`$, thus the dependence of $`T_c`$ on the range $`L`$ may not be quantitatively reliable. However it is quite clear that the approximations we made tend to underestimate the importance of the fluctuation effects; thus the qualitative conclusion that $`T_c0`$ as $`L\mathrm{}`$ must hold. ## VI Summary and discussion In the bulk of the paper, we have taken the attitude that the model we study here, namely a system with long (compared to the coherence length) but finite range pairing interaction, is a theoretical model that is interesting in its own right, and worked out some unusual properties of this model, with emphasis on those properties that are qualitatively different from those of the standard weak coupling BCS superconductors stabilized by a short-range pairing potential. Our most interesting finding is that in this model the transition temperature $`T_c`$ is controlled by thermal fluctuations of collective modes, and can be significantly lower than the quasiparticle gap $`\mathrm{\Delta }`$ or the mean-field transition temperature $`T_c^{MF}`$; as a consequence in the temperature range $`T_c<T<T_c^{MF}`$ the system exhibits pseudogap behavior as the electrons are still paired while there is no superconducting long-range order. In this section we attempt to make contact between our results and the phenomenology of cuprate superconductors, discuss the relation between our model and existing theoretical work on the pseudogap behavior, point out the limitations of our model as well as of our analysis, and indicate some natural extensions and directions for future study. One of the motivations of the present study is the observation that the coherence length $`\xi `$ is much shorter in the cuprates than in conventional superconductors. We are, however, by no means the first to suggest that a short $`\xi `$ can lead to behavior qualitatively different from weak coupling BCS theory. Since the early days of high $`T_c`$, following the work of Leggett and Nozieres and Schmitt-Rink, Randeria and coworkers, as well as others, have argued that the short coherence length may bring the cuprates to a regime that is intermediate between the weak-coupling BCS limit and the Bose-Einstein Condensation (BEC) limit of Cooper pairs. The latter case is realized if $`\xi `$ is much shorter than the inter-particle spacing so that $`k_F\xi 1`$; in terms of energy scales, that corresponds to the case $`\mathrm{\Delta }E_F`$. In this case the Cooper pairs are so closely bound that they hardly overlap, and may be viewed as point-like hard-core bosons at low energies, while the transition temperature $`T_c`$ is essentially their Bose-Einstein condensation temperature which is much lower than and unrelated to the pairing energy $`\mathrm{\Delta }`$. For $`T_c<T<\mathrm{\Delta }`$, the electrons remain paired yet there is no long-range superfluid order, hence pseudogap behavior. In terms of low-energy excitations responsible for destroying superconductivity, in the BEC limit it is the linear Goldstone mode (assuming no Coulomb interaction), while the quasiparticle excitations (broken pairs) cost too much energy to have any effect on $`T_c`$. Put differently, the thermal fluctuations of this Goldstone mode are the classical phase fluctuations of the superconducting order parameter that control the transition in this limit. (We should note that Emery and Kivelson have argued that classical phase fluctuations are the physics of the pseudogap regime in the underdoped materials, but their point of departure is logically distinct, deriving from a small zero temperature superfluid stiffness that is connected to the physics of a doped Mott insulator.) In the cuprates, $`k_F\xi 10`$, and $`\mathrm{\Delta }`$ is still a small energy compared to $`E_F`$ (although the difference is not nearly as overwhelming as in conventional superconductors); we are thus still somewhat distant from the BEC limit. The interesting new feature of the model studied here is that by having $`\xi `$ much smaller than the range of pairing interaction $`L`$, we can get the pseudogap behavior while staying the weak coupling regime (in the sense $`k_F\xi 1`$ and $`\mathrm{\Delta }/E_F1`$). In this case the Goldstone mode is unable to drive $`T_c`$ below the scale of $`\mathrm{\Delta }`$ by itself; it needs all the help from the other collective modes supported by this model. The low-energy spectra of the weak-coupling BCS superconductor, the BEC superconductor, and the present model are summarized schematically in Fig. 1. While there are clearly qualitative differences between the BEC (as well as phase fluctuation) picture and the present model, they share the common spirit that $`T_c`$ is determined by collective modes instead of quasiparticle excitations. It is also worth mentioning that in real systems with long-range Coulomb interaction, the energy of the Goldstone mode is expected to be pushed up to the plasmon frequency; the importance of this fact to the thermal fluctuations of this mode (or classical phase fluctuations) is still under discussion. On the other hand the Coulomb interaction is not expected to significantly affect the spectra of the (gapped) exciton-like modes discussed here. This is due to the different nature of the Goldstone mode and exciton modes: the former is a consequence of the broken gauge symmetry, and introducing the Coulomb interaction (or, more generally, the electromagnetic interaction) converts the Goldstone mechanism of broken symmetry to the Higgs mechanism. The exciton modes, on the other hand, are the bound states formed by quasiparticle pairs due to the residual attractive interaction between quasiparticles not included in the mean field approximation. We therefore believe the presence of Coulomb interaction does not affect the fluctuation physics discussed here significantly. As pointed out already in the Introduction, in some theoretical models for cuprate superconductivity, the range of the interaction that gives rise to Cooper pairing can be very long. In the interlayer pair hopping model, it is assumed that pairing is induced by a pair hopping term in the Hamiltonian, that is diagonal in momentum space. Fourier-transforming to real space, this corresponds to an infinite range hopping term for Cooper pairs, corresponding (loosely) to the $`L\mathrm{}`$ limit of the model we study here (the fact that the hopping term is off-diagonal in layer index is of no qualitative consequence). In that specific form, it has already been shown that the model may be solved exactly in the absence of any other in-plane paring interaction, the transition temperature is zero, and there is pseudogap behavior at low temperatures. Our results are in agreement with this observation. The model we use here, however, is more general and versatile, and in particular enables one to address how the pseudogap behavior develops as the range $`L`$ increases, and how $`T_c`$ approaches zero as $`L\mathrm{}`$. It is equally interesting to scrutinize the results obtained here in the context of the spin fluctuation theory of cuprate superconductivity. In this model the range of the interaction $`L`$ is essentially the spin-spin correlation length. Thus $`L`$ is of order lattice spacing in the overdoped region of the phase diagram, increases as the doping level $`x`$ decreases, becomes much longer than the lattice spacing in the underdoped region, and eventually diverges upon approaching the antiferromagnetic phase boundary at very low doping. The pseudogap behavior is observed in the underdoped regime, where $`T_c0`$ while the gap (the maximum of the $`d`$-wave gap measured at very low $`T`$ by, say, photoemission) slowly increases as $`x`$ decreases. This behavior is certainly consistent with our findings here; the reduction of $`T_c`$ and its departure from the gap is due to the increase of $`L`$, while the size of gap, which in our model saturates at the depth of the potential well $`V_0`$ for large $`L`$, would be set by the scale of the near neighbor spin coupling strength $`J`$. Hence at a very crude level, we can qualitatively account for the phenomenology of the underdoped cuprates by combining our results with the spin-fluctuation model. On the experimental side, some circumstantial evidence for both the importance of the longer range part (beyond near-neighbor) of the pairing interaction, and the possible relevance of our results, exists. i) Recent photoemission measurements of gap anisotropy have found deviations from the standard $$\mathrm{\Delta }_𝐤\mathrm{cos}k_x\mathrm{cos}k_y$$ (120) dependence of the $`d_{x^2y^2}`$ order parameter in the underdoped region, with the deviation increasing with the decreasing doping level. This deviation is interpreted as due to longer-ranged pairing interaction, as a nearest-neighbor attraction leads to Eq. (120). We consider this to be (somewhat indirect) evidence that the range of the pairing interaction increases with decreasing doping level in underdoped cuprates. (ii) It has been recently noticed that there is a strong correlation between $`T_c`$ and peak width of the normal state spin susceptibility $`\chi (q,\omega )`$, in some cuprates. Specifically, Balatsky and and Bourges found that in $`YBCO_{123}`$ and $`La_{214}`$ compounds, $`T_c\delta q`$, where $`\delta q`$ is the width of $`Im\chi (q,\omega )`$ at low $`\omega `$. This lead the authors to conclude that “antiferromagnetism is likely responsible for the high $`T_c`$ superconducting mechanism”. If so, one is lead to the conclusion that $`T_c1/L`$, where $`L1/\delta q`$ is the range of the pairing interaction mediated by the normal state antiferromagnetic spin fluctuations, in agreement with our Eq. (119)! However, given the facts that our model is greatly oversimplified as far as the cuprates go (see below) and our treatment of the fluctuation effects is still preliminary (also below), we consider this agreement to be fortuitous at this point. Nevertheless, the findings of Ref. do indicate the importance of the range of interaction on $`T_c`$, and are in qualitative agreement with our results. (iii) Experimentally, it has been found that upon cooling underdoped cuprate samples, the temperature $`T^{}`$ at which the pseudogap opens up depends on the location in momentum space; $`T^{}`$ is highest near the superconducting gap maxima, while lowest near the gap nodes; it thus suggests that $`T^{}`$ is a “local” property in momentum space. This can be understood quite naturally by invoking a long range pairing interaction as studied here, since such an interaction is localized in momentum space. As we have shown in section III, one can define a momentum-dependent mean field transition temperature, $`T_c^{MF}(ϵ_𝐤)`$, which increases monotonically with $`\mathrm{\Delta }_𝐤`$; it also sets the temperature $`T^{}T_c^{MF}(ϵ_𝐤)`$ below which a local gap is opened up and the pseudogap sets in. We must emphasize that the contact between our model and the cuprate physics made above is tentative, and in its present form this model can only be viewed as an interesting toy model that gives rise to pseudogap behavior. In the following we discuss some limitations of the model as well as our treatment, and some natural directions for extensions. The ground state of our model is a fully-gapped $`s`$-wave superconductor, while the cuprates (at least most of them) are known to have a $`d`$-wave order parameter. The most important difference between the two is that the latter supports gapless nodal quasiparticles. It has already been suggested that the thermally excited quasiparticles may be responsible for emptying the superfluid stiffness, setting the scale of $`T_c`$, and giving rise to pseudogap behavior in underdoped cuprates. This important piece of physics is missing in our model. On the other hand, as long as collective modes of the order parameter are concerned, our method can be generalized to $`d`$-wave (or other unconventional superconductors) fairly easily. The key feature of a long (but finite) range pairing interaction is its locality in momentum space, i.e., $`V_𝐤`$ is sharply peaked in $`𝐤`$ space. This allows a gradient expansion in momentum space for the order parameter. In an $`s`$-wave superconductor, $`V_𝐤`$ is sharply peaked at $`𝐤=0`$, while for a $`d`$-wave superconductor appropriate for cuprates, one would choose a $`V_𝐤`$ that is sharply peaked at a wave vector $`𝐐`$ which is at or near $`(\pi ,\pi )`$. This, however, is not going to make any qualitative difference to the analysis made in this paper. At a more detailed level, two cases need to be distinguished from each other. (i) There is Fermi surface nesting and $`𝐐`$ is exactly or very close to the nesting wave vector. In this case the entire Fermi surface participates in pairing actively and the analysis carried out here, based on the gradient expansion of the superconducting order parameter both along and perpendicular to the Fermi surface, carries through straightforwardly. (ii) There is no Fermi surface nesting, or $`𝐐`$ is not close to the nesting wave vector. In this case only certain “hot spots” at the Fermi surfaces that are connected by $`𝐐`$ participate in the pairing actively. In this case one can still develop a gradient expansion within the hot spots. In either case the qualitative features of our results are expected to be robust. We note that in a recent study of the spin-fermion-hot spot model, it has been conjectured that in the limit of infinite spin-spin correlation length, the transition temperature obtained from solving the Eliashberg equation is only the onset of pseudogap behavior, while the real $`T_c`$ vanishes in that limit. This conjecture is in agreement with our results, as the spin-spin correlation length corresponds to the range of interaction $`L`$ in our model. In the present paper, we have developed a systematic way to study the physics of pseudogap behavior due to large $`L`$, and shown explicitly that $`T_c0`$ as $`L\mathrm{}`$. Our analysis of the thermal fluctuations of the collective modes is based on a Ginzburg-Landau free energy functional, which we derive using a functional integration formalism and an expansion in power series of the magnitude of the order parameter. Strictly speaking this power series expansion is valid only in the vicinity of $`T_c^{MF}`$, where the amplitude of the order parameter just starts to develop; at $`T=T_cT_c^{MF}`$, the amplitude of the order parameter is big and such an expansion is no longer appropriate. What is also missing are the contributions from the fluctuations of components of the order parameter with finite Matsubara frequencies; we neglected them on the ground that we are primary interested effects of thermal fluctuations; however at low $`T`$ these components are important to the physics, especially if one is also interested in the quantum fluctuations of the order parameter. Even with these simplifications, the resultant free energy functional is quite complicated, and we need to introduce approximations in the one-loop calculation of $`T_c`$, like neglecting the anisotropy in the internal (relative) space of the order parameter when the center of mass carries a finite momentum. It is quite possible that one can do a better job in analyzing the fluctuation, especially in the limit of $`L/\xi \mathrm{}`$. It is worth noting, however, that all our approximations tend to underestimate the effects of fluctuations, and thus our basic conclusion that $`T_cT_c^{MF}\mathrm{\Delta }`$ when $`L/\xi 1`$, is clearly valid. Despite the various disclaimers made above, we hope the model and the crossover introduced here will serve as a different paradigm of a non-BCS transition to a superconducting state, which can exhibit pseudogap behavior even without leaving the weak coupling regime. Its relevance to cuprate physics is not completely clear at present, but there are encouraging signs that it is worth further pursuit. ###### Acknowledgements. We have benefited greatly from discussions with and comments from Dan Agterberg, Nick Bonesteel, Elbio Dagotto, Lev Gor’kov, Peter Littlewood, and Doug Scalapino. This work was supported by NSF DMR-9971541, the Fairchild and Sloan Foundations (KY), and NSF DMR-9978074, and the Sloan and Packard Foundations (SLS).
warning/0004/hep-ph0004212.html
ar5iv
text
# I. Introduction ## I. Introduction Nowadays, there are no doubts the model of QCD vacuum as the instanton liquid (IL) is the most practical instrument on the chiral scale of QCD. It provides, as the lattice calculations recently confirmed, not only the theoretical background for describing spontaneous chiral symmetry breaking (SCSB) but is mostly powerful in the phenomenology of the QCD vacuum and in the physics of light quarks while considered to propagate by zero modes arising from instantons. The origin of gluon and chiral condensates turns out in this picture easily understandable and both are quantitatively calculated getting very realistic values defined by $`\mathrm{\Lambda }_{QCD}`$ and parameters of instanton and anti-instanton ensemble, for example, $`i\psi ^{}\psi (250MeV)^3`$. Moreover, the scale for dynamical quark masses, $`M350MeV`$, naturally appears and pion decay constant, $`F_\pi 100MeV`$, is then transparently calculated. Another significant advantage of this approach is that the initial formulation starts basically from the first principles and subsequent approximations being well grounded and reliably controlled are plugged in , . It becomes clear especially in recent years when the impressive progress has been reached in understanding the instanton physics on the lattice . Further we are summarizing several things we have learned thinking of the IL theory and trying to answer the challenging questions. Let us start on that stage of the IL approach when its generating functional has already been factorized as $$𝒵=𝒵_g𝒵_\psi ,$$ where eventually the factor $`𝒵_g`$ provides nontrivial gluon condensate while the fermion part $`𝒵_\psi `$ is responsible to describe the chiral condensate in instanton medium and its excitations. It is usually supposed the functional integral of $`𝒵_g`$ is saturated by the superposition of the pseudo-particle (PP) fields which are the Euclidean solutions of the Yang-Mills equations called the (anti-)instantons $$A_\mu (x)=\underset{i=1}{\overset{N}{}}A_\mu (x;\gamma _i).$$ (1) Here $`A_\mu (x;\gamma _i)`$ denotes the field of a single (anti-)instanton in singular gauge with $`4N_c`$ (for the $`SU(N_c)`$ group) coordinates $`\gamma =(\rho ,z,U)`$ of size $`\rho `$ centred at the coordinate $`z`$ and colour orientation defined by the matrix $`U`$. The nontrivial bloc of corresponding $`N_c\times N_c`$ matrices of PP is a part of potential $$A_\mu (x;\gamma )=\frac{\overline{\eta }_{a\mu \nu }}{g}\frac{y_\nu }{y^2}\frac{\rho ^2}{y^2+\rho ^2}U^{}\tau _aU,y=xz,a=1,2,3,$$ (2) where $`\tau _a`$ are the Pauli matrices, $`\eta `$ is the ’t Hooft symbol , $`g`$ is the coupling constant and for anti-instanton $`\overline{\eta }\eta `$. For the sake of simplicity we do not introduce the distinct symbols for instanton ($`N_+`$) and anti-instanton ($`N_{}`$) and consider topologically neutral IL with $`N_+=N_{}=N/2`$. Formulating the variational principle the practical estimate of $`𝒵_g`$ was found $$𝒵_ge^S$$ with the action of IL defined by the following additive functional <sup>1</sup><sup>1</sup>1 In fact, the additive property results from the supposed homogeneity of vacuum wave function in metric space. Eq. (3) looks like a formula of classical physics although it describes the ground state of quantum instanton ensemble. Intuitively clear, this definition will be still valid even when the wave function is nonhomogeneous with the nonuniformity scale essentially exceeding average instanton size or, precisely speaking, being larger (or of the order) than average size of characteristic saturating field configuration. Then each instanton liquid element of such a distinctive size will provide a partial contribution depending on the current state of IL, see next Section. $$S=𝑑z𝑑\rho n(\rho )s(\rho ).$$ (3) The integration should be performed over the IL volume $`V`$ along with averaging the action per one instanton $$s(\rho )=\beta (\rho )+5\mathrm{ln}(\mathrm{\Lambda }\rho )\mathrm{ln}\stackrel{~}{\beta }^{2N_c}+\beta \xi ^2\rho ^2𝑑\rho _1n(\rho _1)\rho _1^2,$$ (4) weighted with instanton size distribution function $$n(\rho )=Ce^{s(\rho )}=C\rho ^5\stackrel{~}{\beta }^{2N_c}e^{\beta (\rho )\nu \rho ^2/\overline{\rho ^2}},$$ (5) $$\nu =\frac{b4}{2},b=\frac{11N_c2N_f}{3},$$ where $`\overline{\rho ^2}=𝑑\rho \rho ^2n(\rho )/n=\left(\frac{\nu }{\beta \xi ^2n}\right)^{1/2},n=𝑑\rho n(\rho )=N/V`$ and $`N_f`$ is the number of flavours. The constant $`C`$ is defined by the variational maximum principle in the selfconsistent way and $`\beta (\rho )=\frac{8\pi ^2}{g^2}=\mathrm{ln}C_{N_c}b\mathrm{ln}(\mathrm{\Lambda }\rho )(\mathrm{\Lambda }=\mathrm{\Lambda }_{\overline{MS}}=0.92\mathrm{\Lambda }_{P.V.})`$ with constant $`C_{N_c}`$ depending on the renormalization scheme, in particular, here $`C_{N_c}\frac{4.66\mathrm{exp}\left(1.68N_c\right)}{\pi ^2\left(N_c1\right)!\left(N_c2\right)!}`$. The parameters $`\beta =\beta (\overline{\rho })`$ and $`\stackrel{~}{\beta }=\beta +\mathrm{ln}C_{N_c}`$ are fixed at the characteristic scale $`\overline{\rho }`$ (an average instanton size). The constant $`\xi ^2=\frac{27}{4}\frac{N_c}{N_c^21}\pi ^2`$ characterizes, in a sense, the PP interaction and Eqs. (3),(4) and (5) describe the equilibrium state of IL. The minor modification of variational maximum principle (see Appendix) leads to the explicit form of the mean instanton size $`\overline{\rho }\mathrm{\Lambda }=\mathrm{exp}\left\{\frac{2N_c}{2\nu 1}\right\}`$ and, therefore, to the direct definition of the IL parameters unlike the conventional variational principle which allows one to extract those parameters solving numerically the transcendental equation only. The quark fields are considered to be influenced by the certain stochastic ensemble of PPs, Eq. (1), while calculating the quark determinant $$𝒵_\psi D\psi ^{}D\psi e^{S(\psi ,\psi ^{},A)}_A.$$ Besides, dealing with the dilute IL (small characteristic packing fraction parameter $`n\overline{\rho }^4`$) one neglects the correlations between PPs and utilizes the approximation of $`N_c\mathrm{}`$ where the planar diagrams only survive. In addition the fermion field action is approached by the zero modes which are the solutions of the Dirac equation $`i(\widehat{D}(A_\pm )+m)\mathrm{\Phi }_\pm =0`$ in the field of (anti-)instanton $`A_\pm `$, i.e. $$\left[\mathrm{\Phi }_\pm (x)\right]_{ic}=\frac{\rho }{\sqrt{2}\pi |x|(x^2+\rho ^2)^{3/2}}\left[\widehat{x}\frac{1\pm \gamma _5}{2}\right]_{ij}\epsilon _{jd}U_{dc}$$ (6) with the colour $`c,d`$ and the Lorentz $`i,j`$ indices and the antisymmetric tensor $`\epsilon `$. In particular, at $`N_f=1`$ the quark determinant reads $$𝒵_\psi D\psi ^{}D\psi \mathrm{exp}\left\{𝑑x\psi ^{}(x)i\widehat{}\psi (x)\right\}\left(\frac{Y^+}{VR}\right)^{N_+}\left(\frac{Y^{}}{VR}\right)^N_{},$$ (7) $$Y^\pm =i𝑑z𝑑U𝑑\rho n(\rho )/n𝑑x𝑑y\psi ^{}(x)i\widehat{}_x\mathrm{\Phi }_\pm (xz)\mathrm{\Phi }_\pm ^{}(yz)i\widehat{}_y\psi (y),$$ where the factor $`R`$ makes the result dimensionless and is also fixed by the saddle point calculation. Much more accurate result for the Green function of quark in the PP ensemble allows one to calculate the generating functional even beyond the chiral limit and the simple zero mode approximation turns out amazingly fruitful again to develop (quantitatively) the low energy phenomenology of light quarks . Thus, the IL approach at the present stage of its development looks very indicative, well theoretically grounded and reasonably adjusted phenomenologically. The proper form of generating functional obtained and its reasonable parameter dependence provide with enough predictive power and justify, hence, the approximations made. Generally, it leads to conclude the quark feedback upon the instanton background is pretty limp and could be perturbatively incorporated as a small variation of instanton liquid parameters $`\delta n`$ and $`\delta \rho `$ around their equilibrium values of $`n`$ and $`\overline{\rho }`$ being in full analogy with the description of chiral condensate excitations. Indeed, the result of nontrivial calculation of the functional integral (treating the zero quark modes in the fermion determinant substantially from physical view point) takes down to ’encoding’ the IL state with just those two parameters. Moreover, the IL density appears in the approach via the packing fraction parameter $`n\overline{\rho }^4`$ only (clear from dimensional analysis) what means one independent parameter existing in practice. It is just the instanton size. The analysis of the quark and IL interaction is addressed in this paper developing our idea of phononlike excitations of IL resulting from the adiabatic changes of the instanton size. Hence, we describe further the quark feedback dealing with these deformable (anti-)instanton configurations which are the field configurations Eq. (2) characterized by the size $`\rho `$ depending on $`x`$ and $`z`$, i.e. $`\rho \rho (x,z)`$. The paper is organized as follows: in Section II we discuss the modification of quark determinant when the functional integral is saturated by the deformable modes at one flavour. Then in Section III we develop the proper calculation (tadpole approximation) which is based on the saddle point method. Section IV is devoted to the generalization for the multiflavour picture. The paper includes also Appendix where the fault finding reader gets a chance to control the explicit formulae for the IL parameters and to improve our calculations if is able. ## II. Dynamics of phononlike excitations Apparently, the essence of what we discuss here could be illuminated in the following way. Saddle point method calculation of the functional integral implies the treatment of the action extremals which are the solutions of classical field equations. For the case in hands the action $`S[A,\psi ^{},\psi ]`$ is constructed including the gluon fields $`A_\mu `$, (anti-)quarks $`\psi ^{},\psi `$ and extremals which are given by the solutions of the consistent system of the Yang-Mills and Dirac equations. As a trial configuration in the IL theory the superposition of (anti-)instantons which is the approximate solution of the Yang-Mills equations (with no backreaction of the quark fields) and a background field for the Dirac equation is simultaneously considered. We believe it is reasonable to make use the deformable (crumpled) (anti-)instantons $`A_\pm (x;\gamma (x))`$ as the saturating configurations. They just admit of varying the parameters $`\gamma (x)`$ of the Yang-Mills sector of the initial consistent system in order to describe the influence of quark fields in the appropriate variables for the quark determinant. Taking the action in the form $`S[A_\pm (x,\gamma (x)),\psi ^{},\psi ]`$ we would receive the corresponding variational equation for the deformation field $`\gamma (x)`$ which approaches most optimally (as to the action extremum) PP at nonzero quark fields. In the field theory, for example, the monopole scattering , the Abrikosov vortices scattering are treated in a similar way. Actually, the unstationary picture seems more adequate as a specific role of instanton-anti-instanton annihilation channel prompts and an investigation of two-body colour problem teaches . In practice, for IL we avoid the difficulties which come with solving the variational equations if we consider the long wavelength excitations only (with the wave length $`\lambda `$ much larger than the characteristic instanton size $`\overline{\rho }`$). Indeed, it can be done because we are searching the kinetic energy of the deformation fields <sup>2</sup><sup>2</sup>2Then we are allowed to take the slowly changing deformation field beyond the integral while calculating the action of deformed instanton. (the one particle contributions) and consider the pair interaction which develops the contact interaction form being calculated in the adiabatic regime . Let us remember here that deriving Eq. (3) we should average over the instanton positions in a metric space. Clearly, the characteristic size of the domain $`L`$ which has to be taken into account should exceed the mean instanton size $`\overline{\rho }`$. But at the same time it should not be too large because the far ranged elements of IL are not ’causally’ related. The ensemble wave function is expected to be homogeneous (every PP contributes to the functional integral being weighted with a factor proportional to $`1/V,V=L^4`$) on this scale. The characteristic configuration which saturates the functional integral is taken as the superposition Eq. (1) with $`N`$ number of PP in the volume $`V`$. It is easy to understand that because of an additivity of the functional, Eq.(3) describes properly even non-equilibrium states of IL when the distribution function $`n(\rho )`$ does not coincide with the vacuum one and, moreover, allows us to generalize this action for the non-homogeneous liquid when the size of the homogeneity obeys the obvious requirement $`\lambda L>\overline{\rho }`$. In particular, the ensemble of deformable PPs as saturating the functional integral obeys these constraints if the corrections coming with the deformation fields $`g_{\mu \nu }`$ are small and the long wavelength excitations (on the scale of the instanton size $`\overline{\rho }`$, i.e. $`\left|\frac{\rho (x,z)}{x}\right|O(1)`$) of initial instanton fields $`G_{\mu \nu }`$ are mainly taken into consideration. Such a condition of smoothness or the adiabatic change of instanton size dictates, in practice, another essential simplification to define the field in the center of instanton $`\frac{\rho (x,z)}{x}\frac{\rho (x,z)}{x}|_{x=z}`$ as a characteristic deformation field in further calculations. In principle, the correction to the instanton field produced by the quark presence might be calculated in its most general form, if the gluon Green function in the instanton medium is known, as $$a_\mu ^a(x,z)=𝑑\xi D_{\mu \nu }^{ab}(xz,\xi z)J_\nu ^b(\xi z_\psi ),$$ (8) where $`J_\nu ^b`$ is the current of external quark source, $`z_\psi `$ belongs to the region of long wavelength disturbance and, at last, $`D_{\mu \nu }^{ab}`$ is the Green function of PP in the instanton medium. In fact, this Green function (even when considered in the field of one instanton) is not well defined but, seems, for the case in hands we could develop the selfconsistent way to calculate the regularized Green function. The nonsingular propagator behaviour in the soft momentum region is defined by the mass gap of phononlike excitations. Fortunately, the exact form of the Green function occurs to be not of great importance here (we are planning to return to the problem of regularized Green function calculation in the forthcoming publication). In the coordinate space it is peaked around the average PP size being in nonperturbative regime and, hence, integral Eq. (8) is estimated to be $$a_\mu ^a(x,z)\overline{D}_{\mu \nu }^{ab}(xz)\overline{J}_\nu ^b(zz_\psi ).$$ (9) On the other hand, the exact instanton definition in the singular gauge, $$A_\mu ^a(x,z)=\frac{\overline{\eta }_{a\mu \nu }}{g}\frac{}{x_\nu }\mathrm{ln}\left(1+\frac{\rho ^2}{y^2}\right),$$ leads to the following correction to the instanton potential $$a_\mu ^a(x)=H_{\mu \nu }^a(x,z)\frac{\rho }{x_\nu },$$ (10) where $`H_{\mu \nu }^a(x,z)=\frac{\overline{\eta }_{a\mu \nu }}{g}\frac{2\rho (x,z)}{y^2+\rho ^2(x,z)}`$. Confining within the precision accepted here we could take $`\rho (x,z)\overline{\rho }`$, i.e. $`H_{\mu \nu }^a(x,z)=H_{\mu \nu }^a(xz)`$. If we compare now both definitions of the correction we are capable to get immediately for $`\rho _\nu `$ the following equation $$H_{\mu \nu }^a(xz)\frac{\rho (x,z)}{x_\nu }=\overline{D}_{\mu \nu }^{ab}(xz)\overline{J}_\nu ^b(zz_\psi ).$$ In fact, the current $`\overline{J}`$ might be taken constant in the long wavelength approximation. Then neglecting the current gradients one is allowed to change the derivatives to obtain the following estimate of the deformation velocity $`\left|\frac{\rho (x,z)}{x}\right|\left|\frac{\rho \left(z\right)}{z}\right|`$ which looks to be well justified since there is no other fields in the problem at all (in the adiabatic approximation). The contribution of deformed (anti-)instantons to the functional integral (when the corrections coming from the PP deformation fields are absorbed) may be estimated as $$S𝑑z𝑑\rho n(\rho )\left\{\frac{\kappa }{2}\left(\frac{\rho }{z}\right)^2+s(\rho )\right\},$$ where $`\kappa `$ is the kinetic coefficient being derived within the quasiclassical approach. Our estimate of it gives the value of a few single instanton actions $`\kappa c\beta `$ with the coefficient $`c1.5÷6`$ depending quantitatively on the ansatz supposed for the saturating configurations. Although this estimate is not much meaningful because there is no the vital $`\kappa `$ dependence eventually (becomes shortly clear) and the kinetic term could be introduced phenomenologically. Thus, this coefficient should be fixed on a characteristic scale, for example $`\kappa \kappa (\overline{\rho })`$, if we not are planning to be beyond the precision peculiar to the approach. Actually, it means adding the small contribution of kinetic energy type to the action per one instanton only. Such a term results from the scalar field of deformations and affects negligibly the pre-exponential factors of the functional integral. In one’s turn the pre-exponential factors do the negligible influence on the kinetic term as well. The deformation fields induced by the dilations and rotations in the isotopic space result in the singular and trivial kinetic coefficients, respectively . Then to evaluate the mass scale related to those modes one needs the heuristic ideas rooted, apparently, beyond the conventional IL and SCSB. If we strive for being within the approximation developed we should retain the small terms of the second order in deviation from the point of action minimum $`\frac{ds\left(\rho \right)}{d\rho }|_{\rho =\rho _c}=0`$ only expecting the approximate validity of $$s(\rho )s(\overline{\rho })+\frac{s^{(2)}(\overline{\rho })}{2}\phi ^2,$$ (11) where $`s^{(2)}(\overline{\rho })\frac{d^2s\left(\rho \right)}{d\rho ^2}|_{\rho _c}=\frac{4\nu }{\overline{\rho ^2}}`$ and the scalar field $`\phi =\delta \rho =\rho \rho _c\rho \overline{\rho }`$ is the field of deviations from the equilibrium value of $`\rho _c=\overline{\rho }\left(1\frac{1}{2\nu }\right)^{1/2}\overline{\rho }`$. Consequently, the deformation field is described by the following Lagrangian density $$=\frac{n\kappa }{2}\left\{\left(\frac{\phi }{z}\right)^2+M^2\phi ^2\right\}$$ with the mass gap of the phononlike excitations $$M^2=\frac{s^{(2)}(\overline{\rho })}{\kappa }=\frac{4\nu }{\kappa \overline{\rho ^2}}$$ which is estimated for IL with $`N_c=3`$, for example, in the quenched approximation to be $$M1.21\mathrm{\Lambda }$$ if $`c=4,\overline{\rho }\mathrm{\Lambda }0.37,\beta 17.5,n\mathrm{\Lambda }^40.44`$ (for the details see the tables of Appendix). Changing the variables we obtain the gluon part of the generating functional as $$𝒵_g^{^{}}D\phi \left|\frac{\delta A}{\delta \phi \mathrm{}}\right|\mathrm{exp}\left\{\frac{n\kappa }{2}𝑑z\left[\left(\frac{\phi }{z}\right)^2+M^2\phi ^2\right]\right\},$$ with the Jacobian $`\left|\frac{\delta A}{\delta \phi \mathrm{}}\right|`$ corresponding those new variables introduced. Let us notice that the latter looks pretty formal because of incomplete set of the transformations shown. However, in the adiabatic approximation, as was mentioned above, the preexponentional Jacobian contribution to the action <sup>3</sup><sup>3</sup>3Generally, the deformation field $`\rho _\nu `$ and integration variable $`a_\mu ^a`$ (10) are related via the rotation matrix: $`\stackrel{~}{a}_\mu ^a=\mathrm{\Omega }_{ab}\mathrm{\Phi }_{\mu \nu }^b(xz)\rho _\nu `$ and in the long-length wave approximation $`\mathrm{\Phi }`$ might be constant $`\mathrm{\Phi }_{\mu \nu }^b(0)`$($`xz`$). With the rotation matrix spanning the colour field $`a_\mu =\mathrm{\Omega }^1\stackrel{~}{a}_\mu `$ on the fixed axis we can conclude that the vectors $`a_\mu ^z`$ and $`\rho _\nu `$ are, in fact, in one to one correspondence (of course, being within one loop approximation and up to this unessential colour rotation). Thus, the Jacobian contribution turns out to be an unessential constant. being a c-number should be omitted. Analyzing the modifications which arise now in the quark determinant $`𝒵_\psi `$ we take into account the variation of fermion zero modes resulting from the instanton size perturbed $$\mathrm{\Phi }_\pm (xz,\rho )\mathrm{\Phi }_\pm (xz,\rho _c)+\mathrm{\Phi }_\pm ^{(1)}(xz,\rho _c)\delta \rho (x,z),$$ where $`\mathrm{\Phi }_\pm ^{(1)}(u,\rho _c)=\frac{\mathrm{\Phi }_\pm (u,\rho )}{\rho }|_{\rho =\rho _c}`$ and because of the adiabaticity it is valid $`\delta \rho (x,z)`$ $`\delta \rho (z,z)=\phi (z)`$. The additional contributions of scalar field generate the corresponding corrections in the factors of the kernels $`Y^\pm `$ of Eq. (7) which are treated in the linear approximation in $`\phi `$ and taking approximately $`\rho _c=\overline{\rho }`$, i.e. $`i\widehat{}_x\mathrm{\Phi }_\pm (xz,\rho )\mathrm{\Phi }_\pm ^{}(yz,\rho )i\widehat{}_y\mathrm{\Gamma }_\pm (x,y,z,\overline{\rho })+\mathrm{\Gamma }_\pm ^{(1)}(x,y,z,\overline{\rho })\phi (z),`$ (12) here we introduced the notations $$\mathrm{\Gamma }_\pm (x,y,z,\overline{\rho })=i\widehat{}_x\mathrm{\Phi }_\pm (xz,\overline{\rho })\mathrm{\Phi }_\pm ^{}(yz,\overline{\rho })(i\widehat{}_y),$$ $$\mathrm{\Gamma }_\pm ^{(1)}(x,y,z,\overline{\rho })=i\widehat{}_x\mathrm{\Phi }_\pm ^{(1)}(xz,\overline{\rho })\mathrm{\Phi }_\pm ^{}(yz,\overline{\rho })(i\widehat{}_y)+i\widehat{}_x\mathrm{\Phi }_\pm (xz,\overline{\rho })\mathrm{\Phi }_\pm ^{(1)}(yz,\overline{\rho })(i\widehat{}_y)$$ with $`i\widehat{}_y`$ left acting operator (the gradients of scalar field $`\phi `$ are negligible according to the adiabaticity assumption again). It is a simple matter to verify that the right hand side of Eq. (12), being integrated over $`dzdU`$, generates the following kernel (in the momentum space) $$\frac{1}{N_c}\left[(2\pi )^4\delta (kl)\gamma _0(k,k)+\gamma _1(k,l)\phi (kl)\right]$$ with $`\gamma _0(k,k)=G^2(k),G(k)=2\pi \overline{\rho }F(k\overline{\rho }/2),\gamma _1(k,l)=G(k)G^{}(l)+G^{}(k)G(l),G^{}(k)=\frac{dG\left(k\right)}{d\rho }|_{\rho =\overline{\rho }},F(x)=2x[I_0(x)K_1(x)I_1(x)K_0(x)]2I_1(x)K_1(x)`$, where $`I_i,K_i(i=0,1)`$ are the modified Bessel functions. The functional integral of Eq. (7) including the phononlike component may be exponentiated in the momentum space <sup>4</sup><sup>4</sup>4In the metric space we have the nonlocal Lagrangian of the phononlike deformations $`\phi (z)`$ interacting with the quark fields $`\psi ^{},\psi `$, i.e. $``$ $`=`$ $`{\displaystyle 𝑑x\psi ^{}(x)i\widehat{}_x\psi (x)}{\displaystyle 𝑑z\frac{n\kappa }{2}\left\{\left(\frac{\phi }{z}\right)^2+M^2\phi ^2(z)\right\}}+`$ $`+`$ $`{\displaystyle \frac{i\lambda _\pm }{N_c}}{\displaystyle 𝑑x𝑑y𝑑z𝑑U\psi ^{}(x)\{\mathrm{\Gamma }_\pm (x,y,z,\overline{\rho })+\mathrm{\Gamma }_\pm ^{(1)}(x,y,z,\overline{\rho })\phi (z)\}\psi (y)}.`$ The physical meaning of the basic phenomenon behind this Lagrangian seems pretty transparent. The propagation of quark fields through the instanton medium is accompanied by the IL disturbance (the analogy with well known polaron problem embarrasses us strongly in this point). with the auxiliary integration over the $`\lambda `$-parameter (see, for example ) $`𝒵_\psi {\displaystyle }{\displaystyle \frac{d\lambda }{2\pi }}\mathrm{exp}\{N\mathrm{ln}\left({\displaystyle \frac{N}{i\lambda VR}}\right)N\}\times `$ $`\times {\displaystyle }D\psi ^{}D\psi \mathrm{exp}\left\{{\displaystyle }{\displaystyle \frac{dkdl}{(2\pi )^8}}\psi ^{}(k)[(2\pi )^4\delta (kl)(\widehat{k}+{\displaystyle \frac{i\lambda }{N_c}}\gamma _0(k,k))+{\displaystyle \frac{i\lambda }{N_c}}\gamma _1(k,l)\phi (kl)]\psi (l)\right\}`$ (we dropped out the factor normalizing to the free Lagrangian everywhere). It is pertinent to mention here the Diakonov-Petrov result comes to the play precisely if the scalar field is switched off. In order to avoid a lot of the needless coefficients in the further formulae we introduce the dimensionless variables (momenta, masses and vertices) $$\frac{k\overline{\rho }}{2}k,\frac{M\overline{\rho }}{2}M,\gamma _0\overline{\rho }^2\gamma _0,\frac{1}{(n\overline{\rho }^4\kappa )^{1/2}}\gamma _1\overline{\rho }\gamma _1,$$ (13) the fields in turn $$\phi (k)(n\kappa )^{1/2}\overline{\rho }^3\phi (k),\psi (k)\overline{\rho }^{5/2}\psi (k),$$ (14) and eventually for $`\lambda `$ we are using $`\mu =\frac{\lambda \overline{\rho }^3}{2N_c}`$. Then the generating functional takes the following form $`𝒵{\displaystyle }d\mu 𝒵_g^{^{\prime \prime }}{\displaystyle }D\psi ^{}D\psi D\phi \mathrm{exp}\{n\overline{\rho }^4(\mathrm{ln}{\displaystyle \frac{n\overline{\rho }^4}{\mu }}1){\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}{\displaystyle \frac{1}{2}}\phi (k)4[k^2+M^2]\phi (k)\}\times `$ (15) $`\times \mathrm{exp}\left\{{\displaystyle \frac{dkdl}{\pi ^8}\psi ^{}(k)2\left[\pi ^4\delta (kl)(\widehat{k}+i\mu \gamma _0(k,k))+i\mu \gamma _1(k,l)\phi (kl)\right]\psi (l)}\right\},`$ where $`𝒵_g^{^{\prime \prime }}`$ is a part of gluon component of the generating functional which survives after expanding the action per one instanton Eq. (11). The functional obtained describes the IL state influenced by the quarks when all the terms containing the scalar field are collected (see also Appendix). As mentioned above we believe this influence analogous to the back impact of phononlike deformations on the quark determinant does not considerably change the numerical results of the IL and SCSB theory. Noninteracting part of phononlike excitation Lagrangian characterizes the IL reaction on the external long-length wave perturbation and, apparently, is its general feature independent of perturbating field nature. The presence of quark condensate makes hint for the appropriate scheme of approximate calculation of the generating functional $$\psi ^{}\psi \phi =\psi ^{}\psi \phi +\{\psi ^{}\psi \psi ^{}\psi \}\phi .$$ ## III. Tadpole approximation The formal integration over the scalar field leads us to the four fermion interaction and the functional integral can not be calculated exactly. However, due to smallness of scalar field corrections we may find the effective Lagrangian substituting the condensate value in lieu of one of the pairs of quark lines (see Fig. 1a.) $$\psi ^{}(k)\psi (l)\psi ^{}(k)\psi (l)=\pi ^4\delta (kl)TrS(k)$$ where $`S(k)`$ is the quark Green function. In such an approach the diagram with four fermion lines in the lowest order of the perturbation theory in $`\mu `$ is reduced to the two-legs diagram with one tadpole contribution (there are two such contributions because of two possible ways of pairing) $`2(i\mu )^2{\displaystyle \frac{dkdldk^{}dl^{}}{\pi ^{16}}\gamma _1(k,l)\gamma _1(k^{},l^{})\psi ^{}(k)\psi (l)\psi ^{}(k^{})\psi (l^{})\phi (kl)\phi (k^{}l^{})}`$ $`4\mu ^2{\displaystyle \frac{dk}{\pi ^4}\gamma _1(k,k)\psi ^{}(k)\psi (k)\frac{dl}{\pi ^4}\gamma _1(l,l)TrS(l)D(0)},`$ where the natural pairing definition was introduced $$\phi (k)\phi (l)=\pi ^4\delta (k+l)D(k),D(k)=\frac{1}{4(k^2+M^2)}.$$ It is obvious the factors surrounding $`\psi ^{}(k)\psi (k)`$ have a meaning of an additional contribution to the dynamical mass $$m(k)=\mu \gamma _1(k,k)(2i\mu )\frac{dl}{\pi ^4}\gamma _1(l,l)TrS(l)D(0),$$ (16) (the initial mass term contains the factor $`2`$ when the dimensionless variables are utilized, i.e. takes a form $`2im`$.) The contribution of the graph with all the quark lines paired (see Fig. 1b) $$2\mu ^2\left[\frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k)\right]^2\pi ^4\delta (0)D(0)=\frac{\mu ^2}{2}\frac{V}{\overline{\rho }^4}\frac{\kappa }{\nu }\left[\frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k)\right]^2,$$ together with contribution of the graph (Fig. 1c) $$2\mu ^2\frac{dkdl}{\pi ^8}\text{ Tr}\gamma _1(k,l)\gamma _1(l,k)S(k)S(l)D(kl),$$ should be taken into account at the same order of the $`\mu `$ expansion while calculating the saddle point equation. Here we used the natural regularization of the $`\delta `$-function $`\delta (0)=\frac{1}{\pi ^4}\frac{V}{\overline{\rho }^4}`$ in the dimensionless units. Then the quark determinant after integrating over the scalar field reads $`𝒵{\displaystyle }d\mu {\displaystyle }D\psi ^{}D\psi \mathrm{exp}\{n\overline{\rho }^4(\mathrm{ln}{\displaystyle \frac{n\overline{\rho }^4}{\mu }}1)+{\displaystyle \frac{2N_c^2}{n\overline{\rho }^4\nu }}{\displaystyle \frac{V}{\overline{\rho }^4}}\mu ^4c^2(\mu )`$ $`2N_c\mu ^2{\displaystyle \frac{V}{\overline{\rho }^4}}{\displaystyle }{\displaystyle \frac{dkdl}{\pi ^8}}\gamma _1^2(k,l){\displaystyle \frac{(kl)\mathrm{\Gamma }(k)\mathrm{\Gamma }(l)}{(k^2+\mathrm{\Gamma }^2(k))(l^2+\mathrm{\Gamma }^2(l))}}D(kl)+{\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}\psi ^{}(k)2[\widehat{k}+i\mathrm{\Gamma }(k)]\psi (k)\}=`$ $`={\displaystyle }d\mu \mathrm{exp}\{n\overline{\rho }^4(\mathrm{ln}{\displaystyle \frac{n\overline{\rho }^4}{\mu }}1)+{\displaystyle \frac{2N_c^2}{n\overline{\rho }^4\nu }}{\displaystyle \frac{V}{\overline{\rho }^4}}\mu ^4c^2(\mu )`$ (17) $`2N_c\mu ^2{\displaystyle \frac{V}{\overline{\rho }^4}}{\displaystyle }{\displaystyle \frac{dkdl}{\pi ^8}}\gamma _1^2(k,l){\displaystyle \frac{(kl)\mathrm{\Gamma }(k)\mathrm{\Gamma }(l)}{(k^2+\mathrm{\Gamma }^2(k))(l^2+\mathrm{\Gamma }^2(l))}}D(kl)+{\displaystyle \frac{V}{\overline{\rho }^4}}{\displaystyle }{\displaystyle \frac{dk}{\pi ^4}}\text{ Tr}\mathrm{ln}[\widehat{k}+i\mathrm{\Gamma }(k)]\},`$ where the vertex function is defined as $$\mathrm{\Gamma }(k)=\mu \gamma _0(k,k)+m(k),$$ and we introduced the function $`c(\mu )`$ convenient for the practical calculations $$c(\mu )=\frac{i(n\overline{\rho }^4\kappa )^{1/2}}{2\mu N_c}\frac{dk}{\pi ^4}\gamma _1(k,k)\text{ Tr}S(k).$$ As seen from Eq. (III. Tadpole approximation) the Green function of the quark field is selfconsistently defined by the following equation $$2[\widehat{k}+i\mathrm{\Gamma }(k)]S(k)=1.$$ Searching the solution in the form $$S(k)=A(k)\widehat{k}+iB(k),$$ we get $$A(k)=\frac{1}{2}\frac{1}{k^2+\mathrm{\Gamma }^2(k)},B(k)=\frac{1}{2}\frac{\mathrm{\Gamma }(k)}{k^2+\mathrm{\Gamma }^2(k)}.$$ Using Eq. (16) and the definitions of $`\mathrm{\Gamma }(k)`$ and $`B(k)`$ we have the complete integral equation $$\mathrm{\Gamma }(k)=\mu \gamma _0(k,k)+N_c\frac{\kappa }{\nu }\mu ^2\gamma _1(k,k)\frac{dl}{\pi ^4}\gamma _1(l,l)\frac{\mathrm{\Gamma }(l)}{l^2+\mathrm{\Gamma }^2(l)},$$ which drives to have the convenient representation of the solution $$\mathrm{\Gamma }(k)=\mu \gamma _0(k,k)+\frac{N_c}{(n\overline{\rho }^4\kappa )^{1/2}}\frac{\kappa }{\nu }\mu ^3c(\mu )\gamma _1(k,k).$$ What concerns the function $`c(\mu )`$ it is not a great deal to obtain $$c(\mu )=\frac{(n\overline{\rho }^4\kappa )^{1/2}}{\mu }\frac{dk}{\pi ^4}\gamma _1(k,k)\frac{\mathrm{\Gamma }(k)}{k^2+\mathrm{\Gamma }^2(k)},$$ and, therefore, the complete integral equation for the function $`c(\mu )`$ which is shown in Fig. 2 for $`N_f=1`$. Let us underline the $`N_f`$-dependence of the $`c(\mu )`$ function in the interval of $`\mu `$ determined by saddle point value is unessential. Then we easily obtain for the additional contribution to the dynamical quark mass $$m(k)=\frac{N_c}{(n\overline{\rho }^4\kappa )^{1/2}}\frac{\kappa }{\nu }\mu ^3c(\mu )\gamma _1(k,k),$$ (18) and see the cancellation of the kinetic coefficient $`\kappa `$ in $`m`$ if we remember the definition (13). Thus, it means the precise value of the coefficient is unessential as declared. We have the following equation for the saddle point of the functional of Eq. (III. Tadpole approximation) $`{\displaystyle \frac{n\overline{\rho }^4}{\mu }}2N_c{\displaystyle \frac{dk}{\pi ^4}\frac{[\mathrm{\Gamma }^2(k)]_\mu ^{}}{k^2+\mathrm{\Gamma }^2(k)}}+2N_c{\displaystyle \frac{dkdl}{\pi ^8}\left\{\frac{\mu ^2\gamma _1^2(k,l)[(kl)\mathrm{\Gamma }(k)\mathrm{\Gamma }(l)]}{(k^2+\mathrm{\Gamma }^2(k))(l^2+\mathrm{\Gamma }^2(l))}\right\}_\mu ^{}D(kl)}`$ (19) $`{\displaystyle \frac{2N_c^2}{n\overline{\rho }^4\nu }}[\mu ^4c^2(\mu )]_\mu ^{}(n\overline{\rho }^4)_\mu ^{^{}}\mathrm{ln}{\displaystyle \frac{n\overline{\rho }^4}{\mu }}=0,`$ where the prime is attributed to the differentiation in $`\mu `$. Being interested in receiving a closed equation we need to know the derivative $`c^{}(\mu )`$ too. Then the definition of $`c(\mu )`$ above allows us to have $$(1\mu ^2A(\mu ))c^{}(\mu )=2\mu A(\mu )c(\mu )+B(\mu ),$$ where $$A(\mu )=\alpha (\mu )\frac{N_c\kappa }{\nu }\frac{dk}{\pi ^4}\frac{\gamma _1^2(k)}{k^2+\mathrm{\Gamma }^2(k)}\frac{k^2\mathrm{\Gamma }^2(k)}{k^2+\mathrm{\Gamma }^2(k)},$$ $$B=\frac{2(n\overline{\rho }^4\kappa )^{1/2}}{\mu ^2}\frac{dk}{\pi ^4}\frac{\gamma _1(k)\mathrm{\Gamma }^3(k)}{(k^2+\mathrm{\Gamma }^2(k))^2},$$ and $$\alpha (\mu )=1\mu ^2\frac{N_c}{\beta \xi ^2}\frac{\mathrm{\Gamma }(\nu +1/2)}{\nu ^{1/2}\mathrm{\Gamma }(\nu )}\frac{c(\mu )}{n\overline{\rho }^4\left(n\overline{\rho }^4\frac{\nu }{2\beta \xi ^2}\right)}.$$ Thereby the saddle point equation absorbs the effect of the IL parameter modification ($`n(\mu )`$) produced by equilibrium instanton size shift $`\rho _c\overline{\rho }`$ which comes in the leading order from simple tadpole graph $`2i\mu {\displaystyle \frac{dkdl}{\pi ^8}\gamma _1(k,l)(\pi ^4)\delta (kl)TrS(k)\phi (kl)}=\mathrm{\Delta }\phi (0),`$ (20) $`\mathrm{\Delta }=2i\mu {\displaystyle \frac{dk}{\pi ^4}\gamma _1(k,k)TrS(k)}={\displaystyle \frac{4N_c}{(n\overline{\rho }^4\kappa )^{1/2}}}\mu ^2c(\mu ),`$ (remind here $`\phi =\rho \rho _c`$, and $`\phi (0)=𝑑z\phi (z)`$ is the scalar field in momentum representation). In the Table 1 the numerical results (M.S.Z.) are shown for $`N_f=1`$ comparing to those Diakonov and Petrov (D.P.) Table 1. | | D.P. | | | M.S.Z. | | | --- | --- | --- | --- | --- | --- | | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | | $`5.2710^3`$ | $`359`$ | $`(333)^3`$ | $`4.8110^3`$ | $`362`$ | $`(322)^3`$ | The parameters indicated in the Table 1 are the dynamical quark mass $$M(0)=2\mathrm{\Gamma }(0)\left(\frac{1}{\overline{\rho }}\right)[MeV]$$ and the quark condensate $$i\psi ^{}\psi =iTrS(x)|_{x=0}=2N_c\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }(k)}{k^2+\mathrm{\Gamma }^2(k)}\left(\frac{1}{\overline{\rho }}\right)^3[MeV]^3.$$ Through this paper the value of renormalization constant is fixed by $`\mathrm{\Lambda }=280MeV`$. Then the IL parameters are slightly different from their conventional values $`\overline{\rho }(600MeV)^1,\overline{R}(200MeV)^1`$ (see the corresponding tables in Appendix). However, with the minor $`\mathrm{\Lambda }`$ variation the parameters could be optimally fitted. As expected the change of quark condensate is insignificant, the order of several $`MeV`$, what hints the existence of new soft energy scale established by the disturbance which accompanies the quark propagation through the instanton medium. ## IV. Multiflavour approach In order to match the approach developed to phenomenological estimates we need the generalization for $`N_f>1`$. Then the quark determinant becomes , $$𝒵_\psi D\psi ^{}D\psi \mathrm{exp}\left\{𝑑x\underset{f=1}{\overset{N_f}{}}\psi _f^{}(x)i\widehat{}\psi _f(x)\right\}\left(\frac{Y^+}{VR^{N_f}}\right)^{N_+}\left(\frac{Y^{}}{VR^{N_f}}\right)^N_{},$$ $$Y^\pm =i^{N_f}𝑑z𝑑U𝑑\rho n(\rho )/n\underset{f=1}{\overset{N_f}{}}𝑑x_f𝑑y_f\psi _f^{}(x_f)i\widehat{}_{x_f}\mathrm{\Phi }_\pm (x_fz)\mathrm{\Phi }_\pm ^{}(y_fz)i\widehat{}_{y_f}\psi _f(y_f).$$ With phononlike component included every pair of the zero modes $`\mathrm{\Phi }\mathrm{\Phi }^{}`$ acquires the additional term similar to Eq. (12). The appropriate transformation driving the factors $`Y^\pm `$ to their determinant forms is still valid here since the correction term differs from the basic one with the scalar field $`\phi `$. The complete integration over $`dz`$ leads (in the adiabatic approximation $`\phi (x,z)\phi (z)`$) to the transparent Lagrangian form with the momentum conservation of all interacting particles. Besides, we keep the main terms of $`Y^\pm `$ expanding in $`\phi `$. The quark zero modes generate the factor similar to Eq. (II. Dynamics of phononlike excitations) with $`\frac{1}{N_c}`$ being changed by the factor $`\left(\frac{1}{N_c}\right)^{N_f}`$ and then in the leading $`N_c`$ order we have $`Y^\pm =\left({\displaystyle \frac{1}{N_c}}\right)^{N_f}{\displaystyle 𝑑z\underset{N_f}{det}\left(iJ^\pm (z)\right)},`$ $`J_{fg}^\pm (z)={\displaystyle \frac{dkdl}{(2\pi )^8}\left[e^{i(kl)z}\gamma _0(k,l)+\frac{dp}{(2\pi )^4}e^{i(kl+p)z}\gamma _1(k,l)\phi (p)\right]\psi _f^{}(k)\frac{1\pm \gamma _5}{2}\psi _g(l)}.`$ While providing the functional with the Gaussian form we perform the integration over the auxiliary parameter $`\lambda `$ together with the bosonization resulting in the integration over the auxiliary matrix $`N_f\times N_f`$ meson fields $$\mathrm{exp}\left[\lambda det\left(\frac{iJ}{N_c}\right)\right]𝑑\mathrm{exp}\left\{iTr[J](N_f1)\left(\frac{det[N_c]}{\lambda }\right)^{\frac{1}{N_f1}}\right\}.$$ As a result the generating functional may be written as $`𝒵={\displaystyle }{\displaystyle \frac{d\lambda }{2\pi }}𝒵_g^{^{\prime \prime }}\mathrm{exp}(N\mathrm{ln}\lambda ){\displaystyle }D\phi \mathrm{exp}\{{\displaystyle }{\displaystyle \frac{dk}{(2\pi )^4}}{\displaystyle \frac{n\kappa }{2}}\phi (k)[k^2+M^2]\phi (k)\}`$ $`{\displaystyle }D_{L,R}\mathrm{exp}{\displaystyle }dz\{(N_f1)[\left({\displaystyle \frac{det[_LN_c]}{\lambda }}\right)^{\frac{1}{N_f1}}+\left({\displaystyle \frac{det[_RN_c]}{\lambda }}\right)^{\frac{1}{N_f1}}]\}`$ (21) $`{\displaystyle }D\psi ^{}D\psi \mathrm{exp}\{{\displaystyle }{\displaystyle \frac{dk}{(2\pi )^4}}{\displaystyle \underset{f}{}}\psi _f^{}(k)(\widehat{k})\psi _f(k)+i{\displaystyle }dz(Tr[_LJ^+]+Tr[_RJ^{}])\}.`$ Now scalar field interacts with the quarks of the different flavours, nevertheless, the dominant contribution is expected from the tadpole graphs where any pair of the quark fields is taken in the condensate approximation as done at $`N_f=1`$ $$\psi _f^{}(k)\psi _g(l)\psi _f^{}(k)\psi _g(l)=\pi ^4\delta _{fg}\delta (kl)TrS(k).$$ As for the condensate itself we obtain it as the nontrivial solution of saddle point equation. For example, it is $$(_{L,R})_{fg}=\delta _{fg}$$ for the diagonal meson fields. The dimensionless convenient variables (in addition to Eqs. (13), (14)) are the following $$\frac{}{2}\overline{\rho }^3\mu ,\left(\frac{\lambda \overline{\rho }^4}{(2N_c\overline{\rho })^{N_f}}\right)^{\frac{1}{N_f1}}g.$$ Then the effective action ($`𝒵=𝑑g𝑑\mu \mathrm{exp}\{V_{eff}\}`$) in new designations has the form $`V_{\text{eff}}=N(N_f1)\mathrm{ln}g{\displaystyle \frac{V}{\overline{\rho }^4}}(N_f1){\displaystyle \frac{2\mu ^{\frac{N_f}{N_f1}}}{g}}{\displaystyle \frac{V}{\overline{\rho }^4}}{\displaystyle \frac{2N_f^2N_c^2}{n\overline{\rho }^4\nu }}\mu ^4c^2(\mu )+`$ (22) $`+2N_cN_f\mu ^2{\displaystyle \frac{V}{\rho ^4}}{\displaystyle \frac{dkdl}{\pi ^8}\gamma _1^2(k,l)\frac{(kl)\mathrm{\Gamma }(k)\mathrm{\Gamma }(l)}{(k^2+\mathrm{\Gamma }^2(k))(l^2+\mathrm{\Gamma }^2(l))}D(kl)}2N_fN_c{\displaystyle \frac{V}{\overline{\rho }^4}}{\displaystyle \frac{dk}{\pi ^4}\mathrm{ln}\{k^2+\mathrm{\Gamma }^2(k)\}},`$ and the saddle point equation reads $`{\displaystyle \frac{n\overline{\rho }^4}{\mu }}2N_c{\displaystyle \frac{dk}{\pi ^4}\frac{[\mathrm{\Gamma }^2(k)]_\mu ^{}}{k^2+\mathrm{\Gamma }^2(k)}}{\displaystyle \frac{2N_c^2N_f}{n\overline{\rho }^4\nu }}[\mu ^4c^2(\mu )]_\mu ^{}+`$ (23) $`+2N_c{\displaystyle \frac{dkdl}{\pi ^8}\left\{\frac{\mu ^2\gamma _1^2(k,l)[(kl)\mathrm{\Gamma }(k)\mathrm{\Gamma }(l)]}{(k^2+\mathrm{\Gamma }^2(k))(l^2+\mathrm{\Gamma }^2(l))}\right\}_\mu ^{}D(kl)}(n\overline{\rho }^4)_\mu ^{^{}}\mathrm{ln}\left(\left({\displaystyle \frac{n\overline{\rho }^4}{2}}\right)^{1/N_f}{\displaystyle \frac{1}{\mu }}\right)=0.`$ The additional contribution to the dynamical quark mass in Eq. (18) gains the factor $`N_f`$ because the scalar nature of the phononlike field requires to match the tadpole quark field condensates of all $`N_f`$ flavours to each vertex $$m(k)=\frac{N_fN_c}{(n\overline{\rho }^4\kappa )^{1/2}}\frac{\kappa }{\nu }\mu ^3c(\mu )\gamma _1(k,k).$$ Table 2 complements the Table 1 with the calculations at $`N_f=2`$ Table 2. | | D.P. | | | | M.S.Z. | | | | | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`F_\pi `$ | $`F_\pi ^{^{}}`$ | $`\mu `$ | $`M(0)`$ | $`i\psi ^{}\psi `$ | $`F_\pi `$ | $`F_\pi ^{^{}}`$ | | $`4.8310^3`$ | $`386`$ | $`(381)^3`$ | $`135`$ | $`111`$ | $`3.2610^3`$ | $`298`$ | $`(335)^3`$ | $`108`$ | $`90`$ | where $`F_\pi [MeV]`$ is the pion decay constant and $$F_\pi ^2=\frac{N_cN_f}{2}\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }^2(k)\frac{k}{2}\mathrm{\Gamma }^{}(k)\mathrm{\Gamma }(k)+\frac{k^2}{4}(\mathrm{\Gamma }^{}(k))^2}{(k^2+\mathrm{\Gamma }^2(k))^2}\left(\frac{1}{\overline{\rho }}\right)^2,$$ $`F_\pi ^{^{}}[MeV]`$ is its approximated form $$F_\pi ^{}_{}{}^{}2=\frac{N_cN_f}{2}\frac{dk}{\pi ^4}\frac{\mathrm{\Gamma }^2(k)}{(k^2+\mathrm{\Gamma }^2(k))^2}\left(\frac{1}{\overline{\rho }}\right)^2,$$ here $`\mathrm{\Gamma }^{}(k)=\frac{d\mathrm{\Gamma }\left(k\right)}{dk}`$, and the condensate $`i\psi ^{}\psi `$ is implied for the quarks of each flavour. The light particle introduced and which imitates the scalar glueball properties does not affect significantly the SCSB parameters and correctly describes the soft pion excitations of quark condensate. Meanwhile, the experimental status of this light scalar glueball is very vague. We believe the phononlike excitations could manifest themselves being mixed, for instance, with the excitations of the quark condensate in the scalar channel. ## VI. Conclusion In this paper we have developed the consistent approach to describe the interaction of quarks with IL. Theoretically, it is based (and justified) on the particular choice of the configurations saturating the functional integral what is not merely a technical exercise. They are the deformable (crumpled) (anti-)instantons with the variable parameters $`\gamma (x)`$ and in the concrete treatment of this paper we play with the variation of the PP size $`\rho (x,z)`$. In a sense, such an ansatz is strongly motivated by the form of quark determinant which is solely dependent on the average instanton size in the SCSB theory. We have demonstrated that in the long-length wave approximation the variational problem of the deformation field optimization turns into the construction of effective Lagrangian for the scalar phononlike $`\phi `$ and quark fields with the Yukawa interaction. Physically, it allows us to analyze the backreaction of quarks on the instanton vacuum. We have pointed out this influence on the IL parameters as negligible. The modification of the SCSB parameters turns out pretty poor as well. In particular, the scale of quark condensate change amounts to a few $`MeV`$ only. Nevertheless, switching on the phononlike excitations of IL leads to several qualitatively new and interesting effects. The propagation of the quark condensate disturbances over IL happens in this approach to be in close analogy with well-known polaron problem. We imply a necessity to take into account the medium feedback while elementary excitations propagating. Besides, it hints that fitting the parameters $`\overline{\rho },n`$ and renormalization constant $`\mathrm{\Lambda }`$ all together with the alteration of $`s(\rho )`$ profile function we might achieve suitable agreement not only in the order of magnitude. The difficulties which confronted us here illuminate the fundamental problem of gluon field penetration into the vacuum (the instanton vacuum in this particular case) as the most principle one. Indeed, it is a real challenge to answer the question about the strong interaction carrier in the soft momentum region. Perhaps, the light particle of scalar glueball properties which appears inherently in our approach and should manifest itself in the mixture with the excitation of quark condensate in scalar channel ($`\sigma `$-meson) is not bad candidate for that role. By the way, it could be experimentally observed <sup>5</sup><sup>5</sup>5 We receive the effective Lagrangian of Yukawa type which admits the estimate of quark-anti-quark bound state while adapted to the non-relativistic approximation and the inequality $`\frac{\mu \gamma _0\left(0\right)}{4\pi M}\frac{\mu ^2\gamma _1^2\left(0\right)}{n\overline{\rho }^4\kappa }2`$ if valid signals its appearance. In the problem under consideration the left hand side of the inequality is $`O(1)`$. as a wide resonance. Summarizing, we understand our calculation can not pretend to the precise quantitative agreement with experimental data and see many things to be done. We are planning shortly to consider the problem of instanton profile , to make more realistic description of the PP interaction and to push our ansatz beyond the long-length wave approximation analyzing more precisely ’instanton Jacobian’ $`\left|\frac{\delta A}{\delta \phi }\right|`$. The authors have benefited from the discussions with many people but especially with N.O. Agasyan, M.M. Musakhanov, Yu.A. Simonov. The paper was accomplished under the Grants CERN-INTAS 2000-349, NATO 2000-PST.CLG 977482. Two of us (S.V.M. and A.M.S.) acknowledge Prof. M. Namiki and the HUJUKAI Fund for the permanent financial support. ## Appendix The contribution of the quark determinant to the IL action is given by the tadpole diagram Eq. (III. Tadpole approximation) which takes the following form when returned to the dimensional variables (see, Eq. (14)) $$\mathrm{\Delta }\phi =\mathrm{\Delta }\frac{(n\kappa )^{1/2}}{\overline{\rho }^3}\phi (0)=\mathrm{\Delta }(n\overline{\rho }^4\kappa )^{1/2}𝑑\rho \frac{n(\rho )}{n}\frac{dz}{\overline{\rho }^4}\frac{\rho (z)\rho _c}{\overline{\rho }}.$$ Then the IL action, Eq. (3), acquires the additional term $$S=𝑑zn\left\{s\mathrm{\Delta }^{}\frac{\rho \rho _c}{\overline{\rho }}\right\},$$ where $`\mathrm{\Delta }^{}=\frac{4N_c}{n\overline{\rho }^4}\mu ^2c(\mu )`$ and the mean action per one instanton is given by the following functional $`s_1=𝑑\rho s_1(\rho )n(\rho )/n`$ with $$s_1(\rho )=\beta (\rho )+5\mathrm{ln}(\mathrm{\Lambda }\rho )\mathrm{ln}\stackrel{~}{\beta }^{2N_c}+\beta \xi ^2\rho ^2n\overline{\rho ^2}\mathrm{\Delta }^{}(\rho \rho _c)/\overline{\rho }.$$ In order to evaluate the equilibrium parameters of IL we treat the maximum principle $$e^Se^{S_0}e^{SS_0}$$ adapting it to the simplest version (when the approximating functional is trivial $`S_0=0`$). In a sense, this choice of the approximating functional should be a little ’worse’ than in Ref. . Its only advantage comes from the possibility to get the explicit formulae for the IL parameters in lieu of solving the complicated transcendental equation. In equilibrium the instanton size distribution function $`n(\rho )`$ should be dependent on the IL action only, i.e. $`n(\rho )=Ce^{s(\rho )}`$ where $`C`$ is a certain constant. This argument corresponds to the maximum principle of Ref. . Indeed, if one is going to approach the functional (3) as a local form $`s=𝑑\rho s_1(\rho )n(\rho )/n`$ where $`s_1(\rho )=\beta (\rho )+5\mathrm{ln}(\mathrm{\Lambda }\rho )\mathrm{ln}\stackrel{~}{\beta }^{2N_c}+\beta \xi ^2\rho ^2n\overline{\rho ^2}`$, it makes the approach selfconsistent. The functional difference $`ss_1=𝑑\rho \{s(\rho )s_1(\rho )\}e^{s(\rho )}/n`$ being varied over $`s(\rho )`$ leads then to the result $`s(\rho )=s_1(\rho )+const`$ keeping into the mind an arbitrary normalization. The maximum principle results in getting the mean action per one instanton as the IL parameters function, for instance, average instanton size $`\overline{\rho }`$. The corrections generated by the ’shifting’ terms turn out to be small and we consider them in the linear approximation in the deviation $`\mathrm{\Delta }`$. The following schematic expansion exhibits how the major contribution appears $$s_1=\frac{(s+\delta )e^{s\delta }}{e^{s\delta }}\frac{se^s+\delta e^s}{e^s}+\frac{se^s\delta e^ss\delta e^se^s}{e^s^2},$$ (24) here $`\delta (\mathrm{\Delta })`$ stands for a certain small ’shifting’ contribution and $`s`$ is the action generated by the gluon component only. The last term in Eq. (24) is small comparing to the first one and we ignore it. Then it is clear that evaluating the mean action per one instanton is permissible to hold the gluon condensate contribution $`s`$ only (without the ’shifting’ term $`\delta `$) in the exponential. Hence, we have for the mean action per one instanton $`s_1=𝑑\rho s_1(\rho )n_0(\rho )/n_0`$, and $`n_0(\rho )`$ is the distribution function which does not include the ’shifting’ term <sup>6</sup><sup>6</sup>6The ’shifting’ term changes the mass of phononlike excitation insignificantly. The equilibrium instanton size as dictated by the condition $`\frac{ds\left(\rho \right)}{d\rho }|_{\rho =\rho _c}=0`$ is equal to $`\rho _c=(\alpha +\mathrm{\Delta }^{}\beta )\overline{\rho },\alpha =\left(1\frac{1}{2\nu }\right)^{1/2},\beta =\frac{1}{4\nu }\left\{1\alpha \frac{\mathrm{\Gamma }\left(\nu +1/2\right)}{\nu ^{1/2}\mathrm{\Gamma }\left(\nu \right)}\right\}`$ and the second derivative of action in the equilibrium point equals to $`s^{^{\prime \prime }}(\rho _c)=\frac{4\nu }{\overline{\rho ^2}}+\frac{2\nu }{\overline{\rho ^2}}\mathrm{\Delta }^{}\left\{\frac{\mathrm{\Gamma }\left(\nu +1/2\right)}{\nu ^{3/2}\mathrm{\Gamma }\left(\nu \right)}\frac{1}{2\nu \alpha }\right\}`$. Another source of corrections to the kinetic coefficient appears while one considers the instanton profile change $`AA+a`$ where the field of correction is $`a\frac{\rho (x,z)}{x}|_{x=z}`$. This mode could appear within the superposition ansatz Eq. (1) and leads to the modification of quark zero mode ($`D(A+a)\psi =0`$). Fortunately, both corrections to the kinetic term are numerically small.. It is possible to obtain for the average squared instanton size and the IL density that <sup>7</sup><sup>7</sup>7In order not to overload the formulae with the factors making the results dimensionless, which are proportional to the powers of $`\mathrm{\Lambda }`$, we drop them out hoping it does not lead to the misunderstandings. $$r^2\overline{\rho ^2}=\nu \left\{1+\frac{\mathrm{\Delta }^{}}{r\overline{\rho }}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\nu \mathrm{\Gamma }(\nu )}\right\}\nu \left\{1+\mathrm{\Delta }^{}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\nu ^{3/2}\mathrm{\Gamma }(\nu )}\right\},$$ (25) $$n=CC_{N_c}\stackrel{~}{\beta }^{2N_c}\frac{\mathrm{\Gamma }(\nu )}{2r^{2\nu }},$$ (26) where the parameter $`r^2`$ equals to $$r^2=\beta \xi ^2n\overline{\rho ^2}.$$ (27) Expanding $`\mathrm{ln}\rho =\mathrm{ln}\overline{\rho }+\frac{\rho \overline{\rho }}{\overline{\rho }}+\frac{1}{2}\frac{\left(\rho \overline{\rho }\right)^2}{\overline{\rho }^2}+\mathrm{}`$ and using Eq. (25) we can show that $$\frac{𝑑\rho n_0(\rho )\mathrm{ln}\rho }{𝑑\rho n_0(\rho )}=\mathrm{ln}\overline{\rho }+\mathrm{\Phi }_1(\nu ),\frac{𝑑\rho n_0(\rho )\rho }{𝑑\rho n_0(\rho )}=\overline{\rho }+\mathrm{\Phi }_2(\nu ),$$ where $`\mathrm{\Phi }_1,\mathrm{\Phi }_2`$ are the certain function of $`\nu `$ independent of $`\overline{\rho }`$. Besides, the average squared instanton size within the precision accepted obeys the equality $`r^2\overline{\rho ^2}=\mathrm{\Phi }(\nu )`$, and $`\mathrm{\Phi }(\nu )`$ is the function of $`\nu `$ only. Then the mean action per one instanton looks like $$s_1=2N_c\mathrm{ln}\stackrel{~}{\beta }+(2\nu 1)\mathrm{ln}\overline{\rho }+F(\nu )$$ ($`F(\nu )`$ is again the function of $`\nu `$ only and its explicit form is unessential here). Calculating its maximum in $`\overline{\rho }`$ we receive $$\overline{\rho }=\mathrm{exp}\left\{\frac{2N_c}{2\nu 1}\right\},\beta =\frac{2bN_c}{2\nu 1}\mathrm{ln}C_{N_c}.$$ Eqs. (25), (27) allows us to receive the equation for packing fraction parameter $$(n\overline{\rho }^4)^2\frac{\nu }{\beta \xi ^2}n\rho ^4=\frac{\mathrm{\Delta }}{\beta \xi ^2}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\sqrt{\nu }\mathrm{\Gamma }(\nu )}.$$ (28) and for positive root we find $$n=\nu \frac{e^{\frac{8N_c}{2\nu 1}}}{\beta \xi ^2}\left\{1+\mathrm{\Delta }^{}\frac{\mathrm{\Gamma }(\nu +1/2)}{2\nu ^{3/2}\mathrm{\Gamma }(\nu )}\right\},$$ and handling Eq. (26) we determine the constant $`C`$. The IL parameters come about close to the parameter values of the Diakonov-Petrov approach and are shown in the following Table Table 3. | | | D.P. | | | | M.S.Z. | | | --- | --- | --- | --- | --- | --- | --- | --- | | $`N_f`$ | $`\overline{\rho }\mathrm{\Lambda }`$ | $`n/\mathrm{\Lambda }^4`$ | $`\beta `$ | $`N_f`$ | $`\overline{\rho }\mathrm{\Lambda }`$ | $`n/\mathrm{\Lambda }^4`$ | $`\beta `$ | | 0 | 0.37 | 0.44 | 17.48 | 0 | 0.37 | 0.48 | 17.48 | | 1 | 0.30 | 0.81 | 18.86 | 1 | 0.33 | 0.70 | 18.11 | | 2 | 0.24 | 1.59 | 20.12 | 2 | 0.28 | 1.13 | 18.91 | Here $`N_f`$ is the number of flavours ($`N_f=0`$ corresponds to the quenched approximation) and $`N_c=3`$. It is curious to notice the quark influence on the IL equilibrium state provokes the increase of the IL density. In Table 4 we demonstrate the mass gap magnitude $`M`$ and the wave length in the ’temporal’ direction $`\lambda _4=M^1`$. To make it more indicative we show also the average distance between PPs from which it is clear, indeed, that the adiabatic approximation $`\lambda L\overline{R}>\overline{\rho }`$ is valid for the long-length wave excitations of the $`\pi `$-meson type. All the parameters are taken at $`\kappa =4\beta `$ but the primed ones correspond to the kinetic term value of $`\kappa =6\beta `$. Table 4. | $`N_f`$ | $`M\mathrm{\Lambda }^1`$ | $`\lambda \mathrm{\Lambda }`$ | $`M^{^{}}\mathrm{\Lambda }^1`$ | $`\lambda ^{}\mathrm{\Lambda }`$ | $`\overline{R}\mathrm{\Lambda }`$ | | --- | --- | --- | --- | --- | --- | | 0 | 1.21 | 0.83 | 0.99 | 1.01 | 1.2 | | 1 | 1.34 | 0.75 | 1.09 | 0.91 | 1.1 | | 2 | 1.45 | 0.69 | 1.18 | 0.84 | 0.97 | here $`\overline{R}=n^{1/4}`$ is the distance between PPs. | | | | --- | --- | | | | | --- | --- |