id
stringlengths
30
36
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
5
878k
no-problem/9908/hep-lat9908037.html
ar5iv
text
# Ginsparg-Wilson Games ## Abstract I implement a set of tricks for constructing lattice fermion actions which approximately realize the Ginsparg-Wilson relation, with very promising results from simulations. It might be useful to have a simple lattice fermion action $`S=\overline{\psi }D\psi `$ which approximately obeys the Ginsparg-Wilson (GW) relation $$\{\gamma _5,D\}=D\gamma _5RD.$$ (1) Published algorithms cost (apparently) hundreds of times as much as the usual clover action. I describe an approach which costs about a factor of $`6.5\times (N+1)`$ as much as the clover action for an $`N`$th order approximation, and even $`N=1`$ looks quite promising. The ideas in this work are based on three remarkable formulas first published by Neuberger: Introducing a zeroth-order Dirac operator $`D_0`$ and defining $`z=1D_0/r_0`$, a GW action (with $`R=r_0`$) is $$D_{GW}=r_0(1\frac{z}{\sqrt{z^{}z}}).$$ (2) The inverse square root is approximated by $$\frac{1}{\sqrt{z^{}z}}\frac{1}{N}\underset{j=1}{\overset{N}{}}\frac{1}{c_jz^{}z+s_j}$$ (3) ($`c_j=\mathrm{cos}^2(\pi (j+1/2)/(2N))`$, $`s_j=1c_j`$) and $$1\frac{A}{B}=\frac{BA}{B}.$$ (4) Here $`BA=W^{(N)}`$ is the polynomial $$\underset{j=1}{\overset{N}{}}(c_jz^{}z+s_j)(z/N)\underset{j}{}\underset{ij}{}(c_iz^{}z+s_i).$$ (5) To use this for propagators, note $`D_{GW}^{(N)}\psi =\varphi `$ is $$\psi =(D_{GW}^{(N)})^1\varphi =B(W^{(N)})^1\varphi .$$ (6) (i.e. $`\psi `$ is found by inverting the simple differential operator $`W`$, and then multiplying by the local operator $`B`$.) Of course, one needs a $`D_0`$ for which Eqn. 3 works well for small $`N`$. A good $`D_0`$ should already be very chiral. This immediately suggests that we begin with a fat link action–these actions are already quite chiral as shown by their small mass renormalization and $`Z_A1`$) . The eigenvalues of a GW action lie on a circle. I determine the best $`D_0`$ by taking a free field test action and varying its parameterization to optimize its eigenvalue spectrum (in the least-squares sense) for circularity, for some $`r_0`$ (which can also be varied; the optimal value is about 1.6). The action of choice is ”planar:” it has scalar and vector couplings $`S=_{x,r}\overline{\psi }(x)(\lambda (r)+i\gamma _\mu \rho _\mu (r))\psi (x+r)`$ for $`r`$ connecting nearest neighbors ($`\stackrel{}{r}=\pm \widehat{\mu }`$; $`\lambda =\lambda _1=0.170`$, $`\rho _\mu =0.177`$) and diagonal neighbors ($`\stackrel{}{r}=\pm \widehat{\mu }\pm \widehat{\nu }`$, $`\nu \mu `$; $`\lambda =\lambda _2=0.061`$, $`\rho _\mu =0.053`$; $`\lambda (r=0)=8\lambda _124\lambda _2`$). The approach of the eigenvalues to a circle is shown in Fig. 1. The massive action for bare mass $`m`$ is obtained from the $`m=0`$ one by $`D(m)=(1+am/2)D_0+am`$. One might think that the iteration could be done starting with the Wilson or clover action. The trick of Eqn. 3 does rapidly pull the eigenmodes onto a circle, but the problem is the decomposition into the $`W/B`$ form. Unlike for the planar action, the eigenvalues of $`W`$ are thrown far out into the complex plane. This is shown in Fig. 2. Unfortunately, $`W`$ is the matrix which is to be inverted for propagators. Since the high momentum modes of a fat link action don’t see the gauge fields very well, they behave like free field modes. The wide spread of eigenvalues means that in real simulations, $`W^{(N)}`$ becomes ill-conditioned even for small $`N`$. Chiral properties of the action, in four dimensions, are tested first by computing the value of its smallest real eigenvalue $`\lambda `$ on a set of isolated instanton configurations (the instanton radius is $`\rho `$) (Fig. 3). In an exact GW action the real eigenvalue would be zero until the instanton fell through the lattice, when it would disappear. In an ordinary action, $`\lambda `$ is a smooth function of $`\rho `$, close to zero for big $`\rho `$ and moving away from zero, generally to a positive value, as $`\rho `$ decreases, until the eigenvalue collides with a doubler and goes imaginary. For a better action, $`\lambda `$ keeps closer to zero and breaks away more steeply, with a step function for $`\lambda `$ as the desired limiting result. As shown by the pion mass in Fig. 4 (quenched, for $`SU(3)`$, $`a=0.2`$ fm, $`8^3\times 24`$ lattice),the zeroth order action is already very chiral and $`N=1`$ iteration is even more so. In Fig. 5 I show the $`N/\rho `$ mass ratio at $`\pi /\rho =0.7`$ for the $`N=0`$ and 1 versions of this action, along with other actions. Both of the new actions (on improved background gauge configurations) have small scaling violations for this observable. As an added feature, the hadron dispersion relation for these actions is better than for the clover action (the extra terms in the planar action can be tuned to optimize this). Finally, a rough calculation of $`Z_A`$ from Ward identities produces a value quite close to unity–as the fat link clover action gave. To conclude: this is an approach towards the construction of a GW action in which all the cost is “up-front” in the evaluation of $`D_0`$, but the gain is that probably only a few terms (maybe just $`N=0`$ or 1) in the expansion of $`D_{GW}^{(N)}`$ are needed. One also only needs to invert the simple (but messy) differential operator $`W^{(N)}`$ (no inverse inside an inverse is needed). It would be very interesting to tackle the hard lattice problem of $`\overline{\psi }\psi `$ along the lines of Ref. with this approach. This work was supported by the US Department of Energy.
no-problem/9908/cond-mat9908049.html
ar5iv
text
# Oscillatory Exchange Coupling across Cr(1-x)Vx Alloy spacers ## Abstract We have identified the pieces of the Fermi surface responsible for the long period oscillations of magnetic coupling across Cr and Cr<sub>(1-x)</sub>V<sub>x</sub> alloy spacers in metallic multilayers. Analysing experiments and results of KKR-CPA calculations we find that the periods are determined by the extremal wave vectors of the hole pockets centered on the N-point in the Brillouin zone. The discovery that the magnetic coupling between two Fe layers separated by a Chromium layer, of thickness $`L`$(=1,100 Å), oscillates from ferromagnetic to anti-ferromagnetic as function of $`L`$ has lead to a veritable avalanche of observations of the same effect in similar metallic multi-layer systems. The interest in this surprising phenomenon derives from two sources. The first is its relation to the technologically important GMR effect. The second is the wealth of evidence which links the periods of the oscillations, $`P_\nu `$, to the Fermi surfaces of the bulk spacers. In this letter we shall be concern with the second aspect of the problem. Whilst there are a variety of theoretical approaches for describing the above oscillatory magnetic coupling, each leads to the conclusion that one is dealing with a variant of the well known RKKY interaction between two impurity spins where the two point defects (spins) are replaced by two planar defects (the magnetic layers). Moreover, they all predict that the periods of oscillations, $`P_\nu `$, are determined by the extremal spanning wave-vectors, perpendicular to the layers, of the Fermi Surfaces of the bulk spacer metals. Reassuringly, there are now many cases where the calculated callipers vectors of the appropriate Fermi surfaces are in good quantitative agreement with measurements of the corresponding periods. Nevertheless, there is no consensus as to which piece of the Chromium Fermi surface is responsible for the long (18 Å) period of oscillation in the original, and since then much studied, experiments on the Fe/Cr/Fe system . In what follows we shall present conclusive evidence that these oscillations are due to the hole pocket centered on the N-point in the bcc Brillouin zone of bulk paramagnetic Cr. Our arguments will be based on KKR-CPA calculations of the Fermi surfaces of Cr<sub>(1-x)</sub>V<sub>x</sub> alloys and a detailed, theoretical and experimental, investigation of the way the Fermi surface evolves as Vanadium is added to Cr. In short, we show that of all the possible candidate extremal vectors only these spanning the N hole-pockets change with concentration in a manner consistent with the experiments of Parkin et al. A second, equally important point we wish to make, concerns the generic problem of the Fermi surfaces in random alloys of transition metals. As has been emphasised repeatedly, these fundamental features of the metallic state vary rapidly with changes in concentration and undergo interesting electronic topological transformations. Using the Cr-V system as an example, the first which features a non trivial topology, we shall demonstrate that measurements of the oscillatory exchange coupling across an alloy can be used as quantitative probe of subtle features of the Fermi surface. In fact, we suggest that they may provide comparable information to that extracted from results of Angular Correlation of (positron) Annihilation Radiation (ACAR) experiments. To be specific, let us recall the general theoretical result that the exchange coupling between magnetic layers, across a pure metal spacer, is given by $$E=\frac{1}{L^2}\underset{\nu }{}A_\nu \mathrm{cos}(\frac{2\pi }{P_\nu }L+\varphi _\nu )$$ (1) where $`A_\nu `$ is the amplitude, $`\varphi _\nu `$ the phase and $`P_\nu `$ is the period of the $`\nu `$-th oscillatory contribution. As was shown in Ref. and is confirmed by our KKR-CPA calculations, the states near the Fermi Energy in the Cr<sub>(1-x)</sub>V<sub>x</sub> alloy case have sufficiently long lifetimes, as described by the Bloch Spectral Function (BSF) $`A(𝐤,E)`$, that the above formula can be used for these random alloys as well as for pure metals. Evidently, this means that spanning vectors are well defined by the sharp peaks in $`A(𝐤,E)`$. Taking this simplifying circumstance for granted we have carried out standard, self-consistent KKR-CPA calculations for a number of concentrations $`x`$ and determined both the Fermi surfaces and their extremal spanning vectors as described in Ref. . All the theoretical results which feature in the foregoing discussion will refer to these calculations. In fig. 1 we display a three dimensional, schematic picture of the Fermi surface of Chromium. Clearly, for such a complex Fermi surface there are many extremal spanning vectors for each direction of the layer growth. Indeed, Stiles find no less than 6 candidates for explaining the long (18Å) period observed on (110) Fe/Cr/Fe multilayers and the controversy as to which of these, if any, is relevant continues unabated. Of course, a parameter free, reliable calculation of the amplitudes and the phases of the oscillatory exchange coupling could differentiate among these candidates. However, such calculations, at the moment, are still very difficult. Model amplitude calculations have been done for Fe/Cr systems either using the reflection probabilities for Fe/Cr interface or employing a Tight Binding model with spin orbit interaction. Whilst they point towards the conclusion that the long oscillation period for pure Cr spacers originates from the N-centered hole pockets, they are not wholly convincing since the size of the periods calculated agrees with the experiments only roughly. In this letter we wish to advocate a much more effective procedure: Namely, we determine the periods $`P_\nu `$ as a function of the Vanadium concentration $`x`$ by the KKR-CPA calculations mentioned above and compare them with these deduced from measurements of the oscillatory coupling. As will be seen presently, only a small set of interrelated periods show the observed dependence on $`x`$. In fig. 2a we show cuts across the Fermi surfaces of pure Cr and the Cr<sub>0.85</sub>V<sub>0.15</sub> alloy parallel to the growth direction, (110), of the multi-layer system. Evidently, the ‘hole like’ pieces of the Fermi surface increase and the ‘electron-like’ ones decrease as $`x`$ changes from $`0`$ to $`0.15`$. Consequently, some of the spanning vectors, identified by arrows, increase or decrease with $`x`$ depending on whether they span electron or hole like pieces of the Fermi surface. The behaviour of the corresponding periods $`P_\nu `$ with increasing V concentration is depicted in fig. 2b. Evidently, only the periods related to the N hole pockets decrease as in the experiments. Thus we conclude, in agreement with the amplitude model calculations of Stiles and Tsetseris et al, that it is these N pockets that are responsible for the observed long period. To generate further evidence in support of the above conclusions all the different N-related spanning vectors, which can be probed by measurements on samples with (110), (100) and (211) growth directions, have been studied. These investigations are summarized in fig. 3. The Fermi surface cross sections in figs. 3a and 3b are well known and correspond to paramagnetic Cr. Evidently N-pocket is not spherical for pure Cr. According to our KKR-CPA calculations as V is added this pocket shrinks without a change in shape or orientation. As a reflection of this fact, the concentration dependences of the N-pocket related extremal wave vectors N<sub>1</sub>, N<sub>2</sub> etc, shown in fig. 3c, are very similar. Clearly, having eliminated the electron pieces of the Fermi surface, like the ‘lens’ or the ‘knob’ as possible contenders for the role of causing the long period oscillation the new issue becomes: the shape and orientation of the N hole pocket. This is important since agreement between theory and experiments should establish the measurement of the oscillatory coupling as novel and powerful probe of the Fermi surface. However such agreement will require further refinement of the experiments. A hopeful sign in this direction is that although Fullerton et al find the oscillations to have the same periods, and even the same amplitude and phase for all three of the growth directions (100), (110) and (211), Tomaz at al reports small differences. Our predictions for the corresponding periods are N<sub>1</sub>, N<sub>5</sub> and N<sub>9</sub> shown in fig. 3c. Clearly, whilst the differences are small, and hence the periods can be said to be roughly the same, in a more discerning experiment, they should be observable. In conclusion, our calculations on the evolution of the Fermi surface of Cr<sub>(1-x)</sub>V<sub>x</sub> with concentration identifies the N-hole pocket as the source of the long period oscillation of the Exchange Coupling across Cr and Cr<sub>(1-x)</sub>V<sub>x</sub> spacers. The N-hole pocket is fairly isotropic but there are little differences in the spanning vectors in different directions which could be measured by measuring the differences of the oscillation periods for different growth directions. Interestingly, the shape and orientation of the N-hole pockets can also be determined by 2-dimensional ACAR experiments. Clearly, a quantitative agreement between the KKR-CPA theory and the two above, very different, experiments would be a significant step towards establishing the measurement of the above discussed periods as a quantitative probe of topologically complex Fermi surfaces such as those which are host to Electronic Topological Transitions in transition metal alloys. Note Added: After submitting the present letter for publication we became aware of the work of D. Koelling who has drawn similar conclusion to us on the basis of the Rigid Band Model calculations. One of the authors (NNL) was supported by the TMR network on ‘Interface Magnetism’ (contract ERBFMRXCT960089) of the European Union. Also two of us (BG and EB) acknowledge financial support by INFM and the use of the facilities at CECUM (Univ. of Messina).
no-problem/9908/astro-ph9908335.html
ar5iv
text
# Clumps of hydrogenous planetoids as the dark matter of galaxies ## 1 Introduction We discuss an accumulation of observational evidence and theoretical predictions supporting the conclusions that the masses of galaxies within about $`10^{21}`$ m (30 kpc) of their central cores are dominated by Proto-Globular-star-Cluster (PGC) mass ($`10^{35}`$ kg $``$ $`10^5M_{\mathrm{}}`$) clumps of planetary mass ($`10^{24}`$ kg $``$ $`10^6M_{\mathrm{}}`$) objects formed from primordial hydrogen-helium gas soon after its transition from plasma $`\mathrm{300\hspace{0.17em}000}`$ years after the Big Bang. The planetoids are termed “rogue planets” by Schild (1996) and “primordial fog particles” (PFPs) by Gibson (1996). The most convincing observational evidence comes from light curves of gravitationally lensed quasars. A few quasars have galaxies precisely along their lines of sight. The overall mass distributions of such galaxies serve as lenses that distort space and cause the quasar to appear as two or more brightened, twinkling, images. The dominant mass component of the lens galaxy determines the dominant twinkle period (or equivalently, the peak frequency of the microlensing light curve spectrum) of the quasar-image light curves, from Einstein’s gravitational equations, as various objects pass in front of the quasar at their estimated transverse velocity. The smaller the object mass, the shorter the period of brightening or darkening (the twinkle period). Keel (1982) was the first to use such evidence to conclude that the masses of galaxies are not dominated by their stars or by other objects with stellar mass. Refsdal and Stabell (1993) noted, from fluctuations of the four images of the “Einstein cross” quasar Q2237, that the mass of the lensing quasar was possibly dominated by objects as small as $`10^{23}`$ kg ($`10^7M_{\mathrm{}}`$), although without image delay corrections for intrinsic quasar variability this conclusion was considered tentative. From a fifteen year record of observations of both A and B images of the first lensed quasar detected (Q0957+561 A,B), Schild (1996) and Schild and Thomson (1997) determined the dominant twinkling frequency of the difference between the A and B brightness curves, corrected for an estimated $`404\pm 26`$ day time delay of image B (now $`416.3\pm 1.7`$, Pelt et al. (1998)) to eliminate brightness variations intrinsic to the quasar. Schild (1996) concluded from the time-delay corrected brightness difference curve, the microlensing record, that the mass of the lensing galaxy of Q0957 is dominated by component objects with planetary mass. A precise quasar lens time delay is required to distinguish between intrinsic brightness changes of the quasar and microlensing events caused by the galaxy mass components, since typical event times were found to be only 10 to 100 days. The time delay was controversial for a number of years, but is now confirmed by several observers (§2) supporting the Schild (1996) conclusion that the lens galaxy mass is dominated by “rogue planets … likely to be the missing mass”. Although star-microlensing collaborations fail to detect planetoids as the halo missing mass and claim to exclude them (§3), their search focused on larger mass dark matter candidate objects (brown dwarfs). In the following section we present records from three observatories (§2) confirming the existence of rapid quasar microlensing events, and confirming that the events tend to occur in clumps. The missing mass of the galaxy therefore consists of primordial planetoidal objects that have not yet accreted to form stars. Clumping of such objects is to be expected as a consequence of the accretional process, with the intermittency of planetoid number density increasing with the mass range. However, tight clumping or any clumping within the clumps of planetoids poses a measurement problem for their detection by the star-microlensing collaborations (§3), which have not yet detected any planetoidal component in the halo mass of the Milky Way Galaxy, contrary to the quasar-microlensing observations of planetoids in other galaxies. Independently and simultaneously with the Schild (1996) observations and conclusions, Gibson (1996) predicted that the mass of all galaxies should be dominated by globular-cluster-mass clumps of planetoids based on a new (non-Jeans) gravitational condensation theory, discussed in the (§4). Details and refinements of the new theory are given by Gibson and Schild 2000a , Gibson (1999), and Gibson 2000a b. We summarize our conclusions in §5. ## 2 A quasar-microlensing event recorded at three observatories A heightened interest in the Q0957 gravitational lens system resulted from a prediction by Kundic et al. (1995) that a rapid decline in the quasar’s brightness should be seen in Feb.—Mar. 1996 in the second arriving B image, based upon observation of the event in the first arriving A image in December 1994. Because the time delay was still controversial, with values of 1.1-years (Schild (1990), Pelt et al. (1994)) and 1.4-years (Lehar et al. (1992), Press et al. (1992)), it appeared that observations during February and May 1996 would settle the time delay issue. Thus at least 3 observatories undertook monitoring programs to observe the predicted event. At Mount Hopkins, the 15 year monitoring program on the 1.2 m telescope continued, with observations made by scheduled observers on 109 nights. Four observations were made each night with a Kron-Cousins R filter, and the observations averaged together for a published nightly mean brightness value. The quasar brightness was referenced to 5 nearby stars whose brightnesses were checked relative to each other to ensure stability of the magnitude zero point. The data are plotted in Figures 1 and 2 of this report, and data for the first season showing the brightness drop in the first-arriving A component have been published by Schild and Thomson (1997). The Princeton data were obtained by Kundic et al. (1995) using the 3.5m Apache Point telescope with g and r filters on the Gunn photometric system. Their data are not published, but data for the first season were posted on a World-Wide-Web site listed in the Kundic report. We have converted these data to a standard R filter using the relations given in Kent (1985). When we compare the Princeton results to Mt. Hopkins data for the same dates, we find an rms residual of 0.029 mag for component A and 0.18 mag for component B. Unexpectedly, we find the origin of this disagreement not to be so much in random errors as in a systematic drift in the apparent zero points in the course of the observing season. Data for the second observing season have not been presented in tabular form, but a plot of the data posted at the Princeton WWW site has allowed us to compare the results. We have taken the Princeton data plot, separated the two colors of data, and rescaled data for the Mt. Hopkins and Canary Island (Oscoz et al. (1996)) groups to make the comparisons in Figures 1 and 2. The Canary Island data are posted at the WWW site given in the report by Oscoz et al. (1996). They were obtained with the 0.8m telescope using standard R filters and local comparison stars. Because data were obtained in response to the Princeton challenge of Kundic et al. (1995), the Canary Island group reports data for the second season only. We have compared their data with the Mt. Hopkins data with the assumption that any Canary Island datum taken within 24 hours of a Mt. Hopkins observation had agreeing dates, and the rms deviations of the two data sets for our 15 agreeing dates is 0.013 mag for image A and 0.016 mag for B. The error estimates listed at the WWW web site are considerably larger, averaging 0.022 for A and 0.020 for B. Because the Canary Island—Mt. Hopkins comparison must have some error contribution from Mt. Hopkins, it is clear that the posted Canary Island error estimates are too large by a factor of approximately 2. In our plots of the Canary Island data, Figure 2, we have used the original posted error estimates. We show in Figure 1 a comparison of the available photometries for the first observing season. In the upper plot, the Mt. Hopkins data are shown with a magnitude scale and zero point for a standard R filter. The Princeton data have been shown with an arbitrary offset of 0.2 mag. These magnitudes are determined from a transformation from the Princeton Gunn g,r photometric system, using the transformation equations determined by Kent (1985). Error bars are shown strictly according to the estimates of the authors. The data are superimposed in the bottom panel of Figure 1, and the error bars are suppressed for clarity. It may be seen that the data agree about as well as predicted from the errors. One artifact that may be noticed is that there appear to be several points, mostly in the Princeton data, markedly below the mean trend. It is surprising that these discrepant points are in the sense of brightness deficiency, because the two principal error sources, cosmic rays and merging of the two quasar images due to bad seeing effects, both tend to make the images brighter. The A component data in the lower panel will be compared to the observations of image B in Figure 3. In Figure 2 we show 3 data sets in the upper panel, with their associated error bars. However as noted previously, we do not actually have the tabulated data for the Princeton team, and we have scaled the results of Mt. Hopkins and the Canary Island groups to the Princeton data as posted in a plot released by the Princeton team. Thus the plotted magnitudes are on the Gunn photometric system, which differs from standard R by a zero point offset and a color term of 0.15 mag. In other words, R = r - 0.15(g-r) + Const. Since image B varied only from 1.071 to 1.142, a variation of 0.07 mag, we conclude that the scatter introduced into the comparison of r and R magnitudes has a full amplitude of 0.01 magnitudes, or a scatter of at most 0.005 magnitudes around a mean offset. Thus we have simply combined the Princeton r magnitudes with an arbitrary zero point offset in the comparison with the Mt. Hopkins and Canary Island R magnitudes in the bottom panel of Figure 2. We find in Figure 2 (bottom) good evidence that the brightness drop predicted by Kundic et al. (1995) did indeed occur at around Julian Date 2450130. The brightness in the R band did indeed drop almost 0.1 magnitudes, and time delays of 423 days (Oscoz et al. (1997)) and 416 days (Kundic et al. (1997)) are determined. However, a remarkable thing happened at the end of this event, or immediately afterward; a strong microlensing event was observed, principally in the Mt. Hopkins data. The event may be seen as a strong downward spike centered on J.D. 2450151. Although the event was primarily seen in the Mt. Hopkins data, the brightness did not recover to the expected level for another 30 days, and for the remainder of this discussion, we refer to this event as the 3-observatory microlens. A much better perspective on the 3-observatory microlens comes from inspection of Figure 3, where we plot data for both observing seasons combined with the 416-day time delay of Kundic et al. (1997). In this plot the open and filled symbols refer to the first and second observing seasons, exactly as in Figures 1 and 2. We consider that from J.D. 2450151 to 2450180 the data records are sufficiently discrepant to conclude that a microlensing event of 30 to 40 days duration and asymmetrical profile occurred. It is of course possible that more than one event was occurring at this time. It is likely that another event was seen at 2450220 $`\pm `$ 10 days, again seen by 3 observatories. A few other significant discrepancies may be recognized in this fascinating combined data record, and it is not surprising that the time delay has been so difficult to determine because of the influence of this complex pattern of microlensing. On the other hand, with the time delay now measured, the microlensing provides a powerful probe of the mass distribution of objects in the lens galaxy, and perhaps elsewhere (Schild 1996). The mass of the object causing the 35 day duration microlensing event at J.D. 2450151 is $`1\times 10^6M_{\mathrm{}}=2\times 10^{24}`$ kg. We now pose the question of the significance level of the detection of microlensing. We avoid questions of a posteriori statistics by phrasing a test as follows. A dramatic microlensing event was seen covering dates J.D. 2450150-70. During the previous year, a brightness record was obtained that covered the same time interval. If we average and smooth the brightness record for the previous year, at what level of statistical significance can we say each observatory noted a departure in the second year? Posed this way, we can easily determine that each of the three observatories observed a departure attributed to microlensing of at least 10 $`\sigma `$, where the standard deviation $`\sigma `$ has been estimated for the individual data points of each observatory. Thus we conclude that each of three observatories has obtained as at least a 10 $`\sigma `$ result that the second arriving image has brightness departures attributed to microlensing, because they were not seen in the first arriving image. ## 3 Comparison with star microlensing searches The search for MAssive Compact Halo Object (MACHO) particles in the Halo of our Galaxy by microlensing stars of the Large Magellanic Cloud (LMC) has resulted in negative but controversial results. The classic Alcock et al. (1995abcde) papers of the MACHO collaboration report lens masses $`m0.1M_{\mathrm{}}`$, for which event times $`t_{sm}130\sqrt{m/M_{\mathrm{}}}`$ (days) are several months. Star-microlensing events by objects with PFP mass $`10^6M_{\mathrm{}}`$ last only 0.13 days (3 hours), and are therefore difficult to detect in the original program which only obtained a single image frame of each field in a night. The corresponding Q0957 quasar-microlensing time $`t_{qm}3\times 10^4\sqrt{m/M_{\mathrm{}}}`$ (days) is 30 days. Alcock et al. (1996) report that for a limited subsample of their data, where several exposures of rapid succession were considered, the low detection rates indicate non-detection of sufficient mass to make PFP’s the entire mass of a standard spheroidal dark matter Halo. Renault et al. (1998) reach the same conclusion from a more intensive search of a smaller area. The combined MACHO and EROS (Expérience de Recherche d’Objets Sombres) collaborations (Alcock et al. (1998)) focus on small-planetary-mass objects such as PFPs in excluding a population with mass $`M_p=(10^710^3)M_{\mathrm{}}`$ as more than 25% of the missing halo mass within 50 kpc of the Galaxy center (the distance to the LMC), or $`M_p=(3.7\times 10^74.5\times 10^5)M_{\mathrm{}}`$ having more than 10%. Here it is important to recognize that the star microlensing projects have not detected the baryonic dark matter, but have only rejected their own model of the dark matter and its distribution. Their assumption that the objects are homogeneously distributed is most unlikely for a small-mass population such as PFPs which are hydrogenous and primordial, and consequently distributed as a complex array of nested clumps due to their nonlinear gravitational-accretion-cascade for a wide range of mass to form stars. For small $`M_p`$ values, the number density $`n_p`$ is likely to become a lognormal random variable with intermittency factor $`I_p\sigma _{ln[n_p]}^20.5ln[M_{\mathrm{}}/M_p]=8.1`$ (Gibson and Schild 2000b ), where $`\sigma _X^2`$ denotes the variance of random variable $`X`$. For a lognormal random variable, the mean to mode ratio is $`exp[3I_p/2]=1.8\times 10^5`$ for $`I_p=8.1`$. A small number of independent samples of $`n_p`$ gives an estimate of the mode (the most probable value) of a random variable, which is what is estimated by MACHO/EROS $`n_p`$ measurements since the LMC occupies only about 0.04% of the sky and the rapid sampling required for a small object search comprised only about 0.2% of their records. An exclusion of $`10^7M_{\mathrm{}}`$ objects as $`0.1M_{halo}`$ from an estimate of the mode of $`n_p`$ is thus not conclusive, since the mean PFP halo mass could be $`1.8\times 10^4M_{halo}`$ from such measurements even if the closely packed star samples are considered independent. Consequently, we suggest that the Alcock et al. (1998) and Renault et al. (1998) interpretations of MACHO/EROS statistics as an exclusion of planetoids comprising the Halo mass are highly model dependent and as yet inconclusive. ## 4 Theory Many astrophysical and cosmological models of structure formation (Padmanabhan (1993), Peebles (1993), Kolb and Turner (1994), Silk (1989), Weinberg (1972), Rees (1976)) are based on the gravitational instability criterion of Jeans (1902). By this criterion, for a homogeneous gas of density $`\rho `$ and sound speed $`V_S`$, the smallest possible scale of gravitational condensation is $`L_JV_S/(G\rho )^{1/2}`$, where $`G`$ is Newton’s gravitational constant. The validity of Jeans’s theory has been questioned by Gibson (1996, 1997ab, 1998, 1999, 2000ab). Why should the speed of sound $`V_S`$ be relevant to gravitational instability? Why are viscous forces and the inertial-vortex forces of turbulent flows neglected? What about magnetic forces and molecular diffusivity? The new theory shows that fluid mechanically determined Schwarz length scale criteria $`L_{SX}`$ apply rather than $`L_J`$, Gibson (1996). Condensation and void formation are possible at lengths matching the largest of the Schwarz scales $`L_{SX}`$, where $`L_{SX}`$ are derived by balancing gravitational forces with viscous, inertial-vortex, or magnetic forces of the fluid, or (for the super-diffusive non-baryonic dark matter) by balancing the diffusion velocity and gravitational velocity of the density field, Gibson (1999), 2000a. The subscript $`X`$ denotes $`V,T,M`$ or $`D`$, respectively. The Jeans theory and its corollary misconceptions “pressure support” and “thermal support” are most misleading when applied to the hot, quiet, early universe, when $`L_JL_{SV}L_{SV}`$ for the baryonic matter. According to the Gibson 1996-2000 hydro-gravitational theory, the Jeans scale $`L_J`$ was irrelevant at the time of first gravitational structure formation $`t10^{12}`$ s in the plasma epoch when $`V_S=c/3^{1/2}`$ was large and $`L_{SV}L_{ST}ctL_J`$, giving proto-supercluster to proto-galaxy mass objects ($`10^{16}M_{\mathrm{}}`$). Recent cosmic microwave background observations of $`\delta T/T`$ show a sub-horizon spectral peak consistent with these first structures but inconsistent with sonic interpretations of current cosmologies. $`L_J`$ sets the formation size and mass of proto-globular-starclusters (PGCs) for $`t10^{13}`$ s at the beginning of the gas epoch, but $`L_J`$ should not be used as a criterion to exclude the simultaneous formation of smaller gassy hydrogenous planetoids (PFPs) at the viscous and weak turbulence Schwarz scales, where $`L_{SV}L_{ST}L_J`$. Our claim is that an important error of current cosmological models is to assume that the Jeans 1902 criterion may be used under all or any circumstances to exclude the gravitational formation of objects or voids. Although it is generally recognized that either powerful turbulence or strong magnetic forces can prevent star formation at $`L_J`$ scales (Chandrasekhar (1951)); for example, in dense molecular clouds produced by supernovas where $`L_{ST}`$ and $`L_{SM}L_J`$, cases of the early universe where $`L_{SX}L_J`$ have been misinterpreted or overlooked. Jeans neglected viscous and inertial-vortex forces in the conservation of momentum equation with gravity. He assumed the pressure depends only on the density. Either of these unjustified assumptions reduce the fluid mechanical problem to one of gravitational acoustics. Jeans also neglected particle and gravitational diffusive terms in the conservation of mass equation, providing another important source of error. As shown in Gibson (1996), the Jeans assumptions are inadequate to describe the highly nonlinear process of gravitational structure formation. The linearized Euler equation Jeans assumed is rarely reliable in any fluid mechanical context, particularly to describe rapidly expanding flows such as that of the early universe where viscous or buoyancy forces are required to suppress turbulence. Strong turbulence was not present at the time of plasma to neutral gas transition, from measurements of the cosmic microwave background (CMB) radiation that show $`\delta T/T`$ values of only $`10^5`$, three or more orders of magnitude less than values expected if the flow were strongly turbulent. It follows that either the Reynolds number or Froude number or both must have been subcritical in the plasma epoch before $`\mathrm{300\hspace{0.17em}000}`$ years to suppress turbulence to the small levels indicated by the CMB, Gibson 2000a b. Gibson (1999) estimates the photon viscosity of the plasma at the time of first structure formation was $`\mathrm{5\hspace{0.17em}10}^{26}`$ $`\mathrm{m}^2\mathrm{s}^1`$, giving an horizon Reynolds number slightly above critical and a viscous Schwarz scale mass $`\rho L_{SV}^3`$ of $`10^{46}`$ kg, the observed mass of a galaxy supercluster, where $`L_{SV}(\nu \gamma /\rho G)^{1/2}`$ is the viscous Schwarz scale, $`\nu `$ is the kinematic viscosity, $`\gamma 1/t`$ is the rate-of-strain of the fluid, and $`\rho `$ is the density. The time $`t`$ when the horizon mass $`\rho (t)L_H^3`$ increases to $`10^{46}`$ kg is about $`10^{12}`$ s, where $`\rho (t)`$ is derived from Einstein’s equation, Gibson 1997b . As the universe expanded and cooled the average density decreased and the viscous condensation scale increased slowly, giving a decrease in the condensation mass to galactic values of about $`10^{42}`$ kg by the time of plasma-gas transition at $`10^{13}`$ s ($`\mathrm{300\hspace{0.17em}000}`$ years). The formation of proto-supercluster and proto-galaxy structures during the plasma epoch is a source of buoyancy forces and subcritical Froude numbers that can partly explain the CMB indications of suppressed turbulence. Because the gravitational free fall time $`\tau _G(\rho G)^{1/2}`$ at each fragmentation stage in the plasma epoch exceeds the universe age $`t`$, no large increase in density could occur in these structures, although voids could form between them. The baryonic density of about $`10^{17}`$ kg $`\mathrm{m}^3`$ estimated at $`t10^{12}`$ s is close to the average density of globular star clusters, and is taken to be a fossil of this time of first fragmentation in the universe. Because of their relatively small mass and rapidity of formation (a few thousand years) from the primordial gas, the $`L_{SV}`$ scale gas planetoids formed within PGCs are termed “primordial fog particles,” or PFPs. PFP formation represents the first gravitational condensation of the universe, when mass density $`\rho `$ first increased due to gravity after $`\mathrm{300\hspace{0.17em}000}`$ years of decreasing. Previous structure formations in the plasma epoch were fragmentations by void formation because gravitational free fall times $`\tau _G(\rho G)^{1/2}`$ required for densities to significantly increase were longer than the age of the universe at neutralization. Proto-superclusters and proto-galaxies densities have monotonically decreased ever since, from an initial value of $`10^{15}`$ (including the nonbaryonic component) at $`t10^{12}`$ s, to present values of about $`10^{23}`$ and $`10^{21}`$ kg $`\mathrm{m}^3`$, respectively, as the “flat” universe has expanded to its present average density $`\rho 10^{26}`$ kg $`\mathrm{m}^3`$. A thermal-acoustic-gravitational instability occurred at the time of neutral gas formation, causing fragmentation of the proto-galaxy gas blobs at the Jeans scale $`L_J10^4L_{SV}`$, simultaneous to the fragmentation at $`L_{SV}`$ scales to form PFPs. From the ideal gas equation $`p/\rho =RTV_{S}^{}{}_{}{}^{2}`$ it follows that changes in the density $`\rho `$ are exactly compensated by changes in the pressure $`p`$ if the temperature $`T`$ is constant, where $`R`$ is the gas constant of the hydrogen-helium mixture. Thus, radiative heat transfer cannot inhibit void formation at scales $`L_{IC}(RT/\rho G)^{1/2}=L_J`$, where $`L_{IC}`$ is termed the initial condensation scale (Gibson and Schild 2000a ). For the primordial gas temperature $`T=\mathrm{3\hspace{0.17em}000}`$ K and primordial gas density $`\rho 10^{17}`$ kg $`\mathrm{m}^3`$, the initial fragmentation mass was $`\rho L_{IC}^310^{35}`$ kg, the mass of a typical globular cluster of stars, with internal fragmentation into a trillion PFPs. Observations that globular clusters typically contain a million small, ancient, stars have often been cited as evidence of the validity of Jeans’s theory, but we see that this thermal-acoustic-gravitational instability has nothing to do with the linear perturbation stability analysis of Jeans (1902). Instead, the $`L_{IC}`$ scale sets the mass of increasingly isolated, $`10^5M_{\mathrm{}}`$ clouds or clumps of PFPs in the expanding proto-galaxy to that of a proto-globular-cluster, or PGC. Relative motions of the $`10^{12}`$ PFPs within the $`10^5`$ PGCs of a proto-galaxy would have been strongly inhibited by drag forces of their gassy environment during their $`\tau _G`$ condensation periods lasting millions of years. Their accretion to form globular cluster stars must have been equally gentle, as evidenced by a remarkable spherical symmetry which Jeans likened to “fuzzy cricket balls”, compared to the chaotic conditions and geometries of present strongly turbulent star forming regions. Old globular star clusters provide fossil evidence of the high density and weak turbulence existing at their time of formation. Young globular star clusters with the same mass and density, Ashman and Zepf (1998), support the present hypothesis that all luminous globular star clusters form, when tidal forces trigger accretion, from the abundance of dark PGC clumps of hydrogenous planetoids permeating all galaxies. Where are these hundreds of thousands of PGC clumps of PFPs formed in each proto-galaxy at the plasma-gas transition? A full review of the evidence is beyond the scope of the present paper. To summarize the model, most PGCs are apparently intact but dark, and reside where they were formed, in galaxy inner-halos and cores. Some PGCs have collided with each other or have been otherwise agitated to activate a full formation of stars; for example, in the 200 observed globular clusters of the Milky Way or the $`\mathrm{10\hspace{0.17em}000}`$ in M87. Some have been disrupted and their PFPs re-evaporated or dispersed, possibly by repeated encounters to form the galaxy disk and core and the dominant mass component of interstellar medium in these luminous spiral galaxy regions. Some may have formed the so-called super-star clusters detected near galaxy cores, with $`10^4M_{\mathrm{}}`$ mass, many large ($`100M_{\mathrm{}}`$) stars, and huge average densities ($`10^{15}`$ kg $`\mathrm{m}^3`$) requiring a compaction process. What provides the material of construction for the 700 young globular star clusters detected in the bright arms of interacting “antennae” galaxies NGC 4038/4039, Whitmore and Schweizer (1995)? Their typical PGC mass, density, and size suggest the bright young PGs form from dark PGCs brought out of cold storage by tidal forces of the galaxy encounter. Random gas clouds would have random smaller densities and masses. Why have PGC densities remained constant? Because the temperature of the universe fell below the 13 K freezing point of their gasses only after about a billion years, the evolution of PFPs in clusters has been gassy, complex, and collisional. Even after freezing to their present mode as rogue Jovian planets they are much more subject to re-evaporation to gas than stars, with consequent drag forces that damp “virial” velocities, and hence are less likely to be expelled from their PGC clusters by collisionless processes than stars from star clusters as discussed by Binney and Tremaine (1987). Thus, although many questions remain, we have good reason to expect that most of the baryonic matter of a galaxy consists of dark clumps of hydrogenous planetoids, as observed by Schild (1996) and predicted by Gibson (1996). ## 5 Conclusions Quasar-microlensing evidence that lens galaxy masses may be dominated by point-mass objects of planetary mass has been reviewed, and it is found that this possibility is not excluded by the present lack of star-microlensing evidence. The assumption of a uniform-number-density of the hypothetical population of planetoids as the halo dark matter of the galaxy is unjustified and unexpected since such a population should be hydrogenous and primordial as the material of construction of galactic stars that are observed and since any accretion process over a million-fold mass range is likely to produce a highly non-uniform number density distribution for the planetoids within a proto-globular-cluster, even if PGCs were found on the plane of a star-microlensing field. Because quasar images are guided by the mass of the lens galaxy, twinkling at the dominant point-mass-frequency is assured. Clumping of clumps, and subsequent clumping of point masses within the clumps, should cause interference and multiple point-mass-lensing with occasional isolated events, as observed, and render the planetoids functionally invisible to sparse star-microlensing samples, Gibson and Schild 2000b . The Schild (1996) inference of planetoids as the missing matter of the Q0957+561 A,B lensed quasar and his estimated 1.1 year time delay of the images are confirmed by three observatories, and supported by the Gibson (1996) nonlinear-hydrodynamic-gravitational condensation theory that independently predicted this result. Furthermore, the Gibson (1996) theory predicts the planetoids should be sequestered in proto-globular-clusters, which can explain the lack of star-microlensing evidence of their existence, along with the probability of strong internal intermittency and clumping of the planetoids within the PGCs due to their accretional cascade. Figure 1. Data for the A (northern) gravitational lens image, recorded in the October 1994 - June 1995 observing season. In the upper panel, brightness estimates with error bars are shown as triangles for the Schild and Thomson (1997) data, and as circles for the Kundic et al. (1995) data. The R magnitude scale is for the Schild and Thomson data and the Kundic et al. data is arbitrarily offset 0.2 mag to permit comparison. In the lower panel, the data are shown superimposed and without error bars, to show the generally good agreement, especially around the date of the large quasar brightness drop at 2449715. In Figures 1, 2, and 3 the most significant 2 digits of the Julian date have been suppressed for clarity. Figure 2. Data for the B (southern) component recorded in the Nov. 1995$``$June 1996 observing season. In the upper panel, filled squares with error bars are from Oscoz et al. 1996, Triangles are from Schild and Thomson (1997), and circles are from the WWW plot at the site reported by Kundic et al. (1995). In the lower panel, data from the three observatories are plotted without error bars. Generally good agreement is shown in the comparison, and distinct brightness trends are seen in all three data sets. Figure 3. Data from the lower panels of Figures 1 and 2 are shown superimposed with the same symbol definitions as previously, and for a 416 day time delay. It may immediately be seen that there is generally good agreement, and that the second arriving B image (solid symbols) generally follows the pattern of fluctuation exhibited the year before in the first-arriving A image (open symbols). However there are important differences; around Julian Date 2450150-70 a strong brightness drop occurred that had not been seen in the first arriving A image. Similarly, around J.D. 2450220 the records differ systematically by several percent.
no-problem/9908/math9908132.html
ar5iv
text
# Untitled Document This paper, while correct, has been withdrawn, as the main result turned out to be an easy consequence of Edwards’ Theorem. This was belatedly discovered by the authors and later pointed out by an anonymous referee.
no-problem/9908/cond-mat9908308.html
ar5iv
text
# Untitled Document The Angle Resolved Photoemission Pseudogap and Anomalous Background of Underdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> as Evidence for a Fermi Level Band Crossing Jason K. Perry and Jamil Tahir-Kheli First Principles Research, Inc. 8391 Beverly Blvd., Suite #171, Los Angeles, CA 90048 www.firstprinciples.com Submitted to Phys. Rev. Lett. Abstract. We show that the unusual observations of a pseudogap in the normal state of underdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> (BiSCO) using angle resolved photoemission spectroscopy (ARPES) is consistent with a new band structure for the cuprate superconductors in which the $`x^2y^2`$ and $`z^2`$ bands are seen to cross at the Fermi level. Limitations in the experimental method prevent the narrow 3D $`z^2`$ band from being fully resolved, leading instead to a broad background with “stepfunction” character. As a consequence, the Fermi surface is mis-assigned and a pseudogap of approximately d-wave symmetry develops. A requirement of the BCS theory of superconductivity and its extensions is the formation of a non-zero gap at the Fermi energy if and only if the material is in the superconducting state $`(T<T_c)`$. Thus the observation of an approximate $`d_{x^2y^2}`$ gap in the normal state $`(T>T_c)`$ of underdoped Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8+δ</sub> (BiSCO) using angle-resolved photoemission spectroscopy (ARPES) is highly unexpected.<sup>1</sup> Present explanations for this phenomenon speculate that above $`T_c`$ Cooper pairs are formed without long-range order of the wavefunction, producing a pseudogap without superconductivity. Below $`T_c`$, the wavefunction has long-range phase order and superconductivity appears. However a detailed theoretical formulation of Cooper pairing without coherence is lacking. In this paper, we argue the ARPES pseudogap is not the result of Cooper pair formation above $`T_c`$, but arises simply from the crossing of two bands at the Fermi level. While conventional LDA band structure calculations do not predict the existence of two such bands, we demonstrated in a series of articles that correlation, which is well acknowledged to be missing in these calculations, changes the band structure more radically than previously assumed.<sup>2</sup> The new band structure is characterized by a narrow 3D $`Cud_{z^2}/O^{}p_z`$ band ($`z^2`$) which crosses the broader, nearly 2D $`Cud_{x^2y^2}/Op_\sigma `$ band ($`x^2y^2`$) at the Fermi level. Our calculated band structure for optimally doped La<sub>1.85</sub>Sr<sub>0.15</sub>CuO<sub>4</sub> (LaSCO) is shown in Figure 1. Significantly, a degeneracy of the two bands is allowed by symmetry along the $`(0,0)(\pi ,\pi )`$ direction of the Brillouin zone. Elsewhere in the Brillouin zone the bands repel. This observation proves to be the essential ingredient in the Interband Pairing Theory (IBP) of high temperature superconductivity.<sup>3</sup> This theory postulates the formation of a new type of Cooper pair (interband pair) comprised of a $`k`$ electron from one band and a $`k`$ electron from another band in the vicinity of the $`x^2y^2`$ and $`z^2`$ symmetry allowed crossing. As described in the above references, IBP can explain a broad range of experimental data from simple band structure arguments. Such experimental data includes the observed d-wave Josephson tunneling (and by extension the ARPES d-wave superconducting gap), the temperature dependence of the Hall effect, the NMR, the mid-IR absorption, and the incommensurate neutron scattering.<sup>3</sup> The origin of the pseudogap in our scenario is due to the difference in the 3D dispersion of the $`x^2y^2`$ and $`z^2`$ bands and the fact that ARPES is a method that maps a Fermi surface in 2D. Electrons ejected from $`k`$ states with predominantly $`z^2`$ character produce a broad linewidth in ARPES and hence an unresolvable quasiparticle peak. A mis-assignment of the Fermi surface results and with it the pseudogap. This pseudogap is a direct measure of $`x^2y^2/z^2`$ band repulsion, which has approximately d-wave symmetry. From these considerations, described in detail below, we conclude the pseudogap is evidence for a Fermi level band crossing and is unrelated to the d-wave superconducting gap. We further show by calculation that the anomalous background spectrum present in all cuprate ARPES data is due to both primary and inelastically scattered secondary electrons associated with the narrow $`z^2`$ band. ARPES on optimally doped and overdoped BiSCO yields a single holelike Fermi surface closed around $`(\pi ,\pi )`$.<sup>4</sup> At temperatures below $`T_c`$, the Fermi surface exhibits a d-wave gap with a node along the $`(0,0)(\pi ,\pi )`$ direction.<sup>5</sup> This gap disappears isotropically as the temperature is increased to $`T_c`$, as one would expect for a superconducting gap. For underdoped BiSCO, the situation is different as follows: 1.) The gap does not appear to be purely d-wave.<sup>6</sup> 2.) The temperature dependence of the gap is anisotropic.<sup>7</sup> 3.) The gap grows larger as the material is further underdoped and $`T_c`$ is decreased. 4.) The gap persists through temperatures above $`T_c`$.<sup>1</sup> Based on our calculated band structure for LaSCO, we conclude the following: 1.) The $`x^2y^2`$ band is broadly dispersing in the CuO<sub>2</sub> planes ($`x`$ and $`y`$ directions) and narrowly dispersing normal to the planes ($`z`$ direction). It is an approximately 2D band. 2.) The $`z^2`$ band is narrowly dispersing in the $`x`$ and $`y`$ directions and moderately dispersing in the $`z`$ direction. It must be considered a 3D band. 3.) The two bands cross at or near the Fermi level. The crossing is allowed by symmetry along the $`(0,0)(\pi ,\pi )`$ diagonal but avoided elsewhere in the Brillouin zone. In order to reconcile this proposed band structure with the Fermi surface mapped by ARPES, one must consider the limitations in the ARPES method.<sup>8</sup> The experimentally measured ARPES linewidth $`\mathrm{\Gamma }_m`$, is a combination of the linewidth or lifetime of the photohole $`\mathrm{\Gamma }_{hole}`$ (hole state left behind by the excited electron) and the linewidth of the photoelectron $`\mathrm{\Gamma }_{elec}`$ (excited state of electron after photon absorption). It is given approximately by, $$\mathrm{\Gamma }_m=\mathrm{\Gamma }_{hole}+(v_{i,z}/v_{f,z})\mathrm{\Gamma }_{elec},$$ $`(1)`$ where $`v_{i,z}`$ is the average Fermi velocity of the photohole in the $`z`$ direction, and $`v_{f,z}`$ is the average Fermi velocity of the photoelectron in the $`z`$ direction. While $`\mathrm{\Gamma }_{hole}0`$ at the Fermi level, $`\mathrm{\Gamma }_{elec}`$ is generally greater than or on the order of $`1.0`$ eV. For the very 2D $`x^2y^2`$ band, dispersion in the $`z`$ direction is negligible compared to that in the $`x`$ and $`y`$ directions. Hence, $`v_{i,z}`$ is expected to be small relative to $`v_{f,z}`$. Thus $`\mathrm{\Gamma }_m\mathrm{\Gamma }_{hole}`$, leading to a resolvable quasiparticle peak and a well defined Fermi surface crossing for $`k`$ states with predominantly $`x^2y^2`$ character. In contrast, the $`z^2`$ band is narrowly dispersing in the $`x`$ and $`y`$ directions but has larger dispersion in the $`z`$ direction. For this band, there is no reason to expect that $`v_{i,z}`$ is not comparable to $`v_{f,z}`$. Thus, the contribution from the linewidth of the photoelectron cannot be neglected, leading to a broad peak cutoff by the Fermi function. For $`k`$ states with predominantly $`z^2`$ character, there will be no resolvable quasiparticle peak but instead a signal that looks like a step function. Given that only the $`x^2y^2`$ band leads to a well resolvable peak with ARPES, the band structure for LaSCO produces a Fermi surface shown in Figure 2, in excellent agreement with recent observations for this material.<sup>9</sup> The $`z^2`$ band contributes only to a broad background signal. Such a signal has been a signature in ARPES on the cuprates, but has always been regarded with confusion.<sup>4-8</sup> Careful experiments with light polarization and photon energy dependence may resolve if this background is due to $`z^2`$ character. Such experiments have been done to confirm that the major resolvable peak is in fact due to $`x^2y^2`$ character.<sup>8,10</sup> The analysis is unfortunately complicated by secondary inelastic scattering processes which may indeed dominate the background signal. Such processes have previously been considered and dismissed as the source of the background because the conventional band structure could not produce a large enough signal and the observed step function character.<sup>8,11</sup> However, in Figure 3, we show the signal due to inelastic scattering from our calculated band structure for LaSCO leads to a background which is significantly larger than that obtained from the conventional $`x^2y^2`$ band structure. The inelastic scattering is proportional to the integrated density of states. Thus, the signal is dominated by the very narrow $`z^2`$ band with a large density of states within 0.1 eV of the Fermi level leading to the correct step function character near the Fermi level. The final ARPES lineshape is composed of primaries from both the $`x^2y^2`$ and $`z^2`$ bands as well as secondaries. The observation of a pseudogap in the cuprates occurs when the true Fermi surface is dominated by $`z^2`$ character. The locus of $`(k_x,k_y)`$ points that comprise the Fermi surface is determined by measuring spectra as one scans through the Brillouin zone from occupied $`k`$ states to unoccupied $`k`$ states. A quasiparticle peak appears as one approaches the Fermi surface and disappears as one scans through the Fermi surface. With our band structure, the dispersing quasiparticle peak will collapse in $`k`$ space when we cross from states with predominantly $`x^2y^2`$ character ($`i.e.`$ narrow linewidth) to states with predominantly $`z^2`$ character ($`i.e.`$ broad linewidth). The extent (and possibly position) of the collapse of the quasiparticle peak is also subject to matrix element effects, which are different for $`x^2y^2`$ and $`z^2`$. Should this crossover occur for $`k`$ states below the Fermi level, the Fermi surface will be mis-assigned. For lower temperatures, the leading edge will be below the Fermi level, leading to the appearance of a pseudogap. At sufficiently high temperatures the $`x^2y^2`$ hole linewidth will be large enough that the leading edge of the spectra will be at the Fermi level, closing the pseudogap anisotropically with temperature.<sup>7</sup> In order to explain the pseudogap in BiSCO, we need to qualitatively understand its band structure. While we have so far only calculated the band structure for LaSCO, simple topological arguments can be used to understand BiSCO. The principal difference between the two materials is that BiSCO has two CuO<sub>2</sub> planes per unit cell instead of one. This means there will be bonding and antibonding combinations of both the $`x^2y^2`$ and $`z^2`$ bands leading to a total of four key bands instead of two. Of these, the two $`x^2y^2`$ bands are nearly degenerate (there is little $`z`$ axis coupling between them), but the two $`z^2`$ bands should be reasonably split in energy such that only three bands (the two $`x^2y^2`$ bands and the antibonding $`z^2`$ band) are important at the Fermi level. The fourth bonding $`z^2`$ band should be lower in energy. In Figure 4, we present a schematic of the dispersion of the three key bands along the symmetry lines $`(0,0)(\pi ,\pi )`$ and $`(\pi ,0)(\pi ,\pi )`$. The shaded region in the figure for $`z^2`$ antibonding shows the spread in the dispersion of this band as a function of $`k_z`$. The two $`x^2y^2`$ bands will vary little as a function of $`k_z`$ due to their approximately 2D character. Along $`(0,0)(\pi ,\pi )`$, the different reflection symmetries of $`x^2y^2`$ versus $`z^2`$ allow the bands to cross. This crossing, which is crucial to IBP, can persist at the Fermi level over a range of dopings due to the $`z`$-axis dispersion of the $`z^2`$ band. From $`(\pi ,0)(\pi ,\pi )`$, only the $`x^2y^2`$ bonding and $`z^2`$ antibonding bands can cross and only if $`k_z=\pi /c`$ or $`k_z=0`$. Elsewhere there is no symmetry to forbid mixing and the three bands must repel. This repulsion has approximately d-wave symmetry. Figure 5 shows the approximate characters of the three bands in a 2D Brillouin zone. One can see that the top band has $`x^2y^2`$ antibonding character near $`(\pi ,\pi )`$ that changes over to $`z^2`$ antibonding character at $`(0,0)`$ and $`(\pi ,0)`$. The middle band has $`x^2y^2`$ bonding character at $`(\pi ,\pi )`$ that changes over to $`z^2`$ antibonding for a small region and then changes again to $`x^2y^2`$ antibonding character at $`(0,0)`$ and $`(\pi ,0)`$. The lowest band is $`z^2`$ antibonding at $`(\pi ,\pi )`$ that becomes $`x^2y^2`$ bonding at $`(0,0)`$ and $`(\pi ,0)`$. For all relevant dopings of BiSCO, the Fermi surfaces of all three bands will be dominated by $`z^2`$ character. Following our argument above, ARPES will mis-assign the Fermi surface as the locus of $`(k_x,k_y)`$ points where character changes from antibonding $`z^2`$ to antibonding $`x^2y^2`$ in Band 2, as indicated by the dashed line. Along the $`(0,0)(\pi ,\pi )`$ diagonal, there are always $`x^2y^2`$ states at the Fermi level since the $`x^2y^2`$ and $`z^2`$ bands can be degenerate here. This leads to a zero gap (i.e. node) along the diagonal. Scanning from $`(\pi ,0)(\pi ,\pi )`$, the $`x^2y^2`$ states lie below the Fermi level due to band repulsion, producing a gap. Thus, the approximately d-wave pseudogap in BiSCO and related materials is due to a simple mis-assignment of the Fermi surface. In regard to the relationship between the pseudogap and the superconducting gap, as argued elsewhere,<sup>3</sup> the symmetry of interband pair to BCS pair scattering produces a d-wave superconducting gap which forces conventional BCS scattering to adopt this phase. The possibility of interband pair to interband pair scattering should produce an additional gap at the nodes, but this would be extremely difficult to observe with ARPES due to its strong $`k_z`$ dependence. The confusion as to the pseudogap arises because when the material is underdoped, the pseudogap is larger in magnitude than the superconducting gap, completely obscuring its presence. As doping is increased, the pseudogap is expected to decrease in magnitude as the Fermi level nears the region where Band 2 switches to $`x^2y^2`$ antibonding character. Simultaneously, the superconducting gap is expected to increase in magnitude. At some point the superconduting gap is expected to be greater in magnitude than the pseudogap, thus obscuring its presence. Eventually the pseudogap will disappear entirely as the Fermi surface becomes increasingly $`x^2y^2`$-like. This behavior is consistent with that observed for underdoped, optimally doped, and overdoped BiSCO. It is important to note that since bands change character smoothly, we have not defined exactly what $`k`$ point is the crossover from $`x^2y^2`$ to $`z^2`$. The crossover momentum is dependent on the sizes of the $`x^2y^2`$ and $`z^2`$ ejection matrix elements and these clearly are dependent upon the incident photon energy. Thus, the controversy over recent observations of a different Fermi surface for BiSCO when the photon energy is $`33`$ eV<sup>12</sup> are not in direct contradiction with the older $`2025`$ eV results, but instead demonstrate that measuring Fermi surfaces using ARPES is not completely straightforward when the relevant bands include orbitals with real 3D dispersion. The authors wish to thank D.S. Dessau, Y.-D. Chuang, and A.D. Gromko for many stimulating discussions. References <sup>1</sup>A.G. Loesser, et al., Science 273, 325 (1996); H. Ding, et al., Nature 382, 51 (1996). <sup>2</sup>J.K. Perry and J. Tahir-Kheli, Phys. Rev. B 58, 12323 (1998); J.K. Perry, J. Phys. Chem., in press (cond-mat/9903088); J.K. Perry and J. Tahir-Kheli, Phys. Rev. Lett., submitted (cond-mat/9907332). See also www.firstprinciples.com. <sup>3</sup>J. Tahir-Kheli, Phys. Rev. B, 58, 12307 (1998); J. Tahir-Kheli, J. Phys. Chem., in press (cond-mat/9903105); J. Tahir-Kheli, Phys. Rev. Lett., to be submitted; J. Tahir-Kheli and J.K. Perry, to be published. See also www.firstprinciples.com. <sup>4</sup>H. Ding, et al., Phys. Rev. Lett. 76, 1533 (1996). <sup>5</sup>Z.-X. Shen, et al., Phys. Rev. Lett. 70, 1553 (1993); H. Ding, et al., Phys. Rev. B 54, 9678 (1996). <sup>6</sup>J. Mesot, et al., to be published (xxx.lanl.gov/cond-mat/9812377). <sup>7</sup>M.R. Norman, et al., Nature 392, 157 (1998). <sup>8</sup>N.V. Smith, P. Thiry, and Y. Petroff, Phys. Rev. B 47, 15476 (1993); Z.-X. Shen and D.S. Dessau, Phys. Rep. 253, 2 (1995). <sup>9</sup>A. Ino, et al., J. Phys. Soc. Japan 68, 1496 (1999). <sup>10</sup>H.Ding, et al., Phys. Rev. Lett. 76, 1533 (1996). <sup>11</sup>L.Z.Liu, J. Phys. Chem. Solids 52, 1471 (1991). <sup>12</sup>Y.-D. Chuang, et al., Phys. Rev. Lett. 83, 3717 (1999); D.L. Feng, et al., to be published (cond-mat/9908056); H.M. Fretwell, et al., to be published (cond-mat/9910221); J. Mesot, et al., to be published (cond-mat/9910430). Figure Captions. Figure 1. Calculated 3D band structure for optimally doped LaSCO (see reference 2). Band dispersion along $`(k_x,k_y)`$ symmetry lines is given for $`k_z=0`$, $`\pi /c`$, and $`2\pi /c`$. Note the $`x^2y^2`$ and $`z^2`$ bands cross along the $`(0,0)(\pi ,\pi )`$ symmetry line but repel near $`(\pi ,0)`$. Figure 2. Calculated 2D Fermi surface (solid line) for optimally doped LaSCO that would be oberved by ARPES. The true Fermi surface also contains 3D character, cross sections of which are represented by the dotted lines. This 3D component contributes to a broad background signal with no resolvable Fermi surface in the ARPES spectrum. Figure 3. Integration over the occupied density of states for our calculated band structure for LaSCO vs. a conventional ($`x^2y^2`$ only) band structure. The ARPES background signal due to inelastic scattering is directly proportional to this curve. The background signal predicted from our band structure is in excellent agreement with that which is observed. Figure 4. Schematic of the dispersion in BiSCO for the $`x^2y^2`$ bonding, $`x^2y^2`$ antibonding, and $`z^2`$ antibonding bands along the $`(0,0)(\pi ,\pi )`$ and $`(\pi ,0)(\pi ,\pi )`$ symmetry lines. The two $`x^2y^2`$ bands cross the $`z^2`$ band along the $`(0,0)(\pi ,\pi )`$ direction, but band repulsion opens up an energy gap in the $`x^2y^2`$ states along the $`(\pi ,0)(\pi ,\pi )`$ direction. The energy scale of the figure is $``$ 0.2-0.3 eV. Figure 5. Schematic of the character of the three key bands in BiSCO. The true Fermi surface (which is approximately indicated by the solid and dotted lines) has significant 3D character and cannot be easily pinned down. As a result, ARPES mis-assigns the Fermi surface as indicated by the dashed line. An approximate d-wave pseudogap is produced from this mis-assignment.
no-problem/9908/hep-ph9908500.html
ar5iv
text
# 1 Introduction ## 1 Introduction A series of polarized lepton-nucleon scattering experiments at CERN, SLAC and DESY has considerably improved our knowledge on the spin structure of the nucleon over the past few years. These experiments were however largely restricted to the inclusive structure function $`g_1(x,Q^2)`$, probing a particular combination of quark polarizations in the nucleon. As a consequence, our picture of the nucleon spin structure is still far from complete. Several future experiments are now attempting to resolve remaining open issues. The spin physics sessions of this workshop made a critical review of the physics potential of these future experiments, with particular emphasis on the potential impact of polarized electron-proton collisions at HERA. With several new and ongoing spin experiments, much information on the proton spin structure will become available in the next few years: complementary measurements at HERMES, COMPASS and RHIC will yield first information on the gluon contribution to the proton spin, combining future results from these experiments with existing structure function data, it will moreover be possible to separate quark and antiquark contributions to the proton spin and to carry out a flavour decomposition. These experiments, as well as future measurements at Jefferson Laboratory will also give first insight into potential higher twist contributions to the proton spin. The kinematical scope of these experiments is however limited, and they will still leave many questions unanswered. Given the numerous new insights HERA has provided into the unpolarized proton structure in the past, the option of polarizing its beams to study proton spin structure appears very tempting. Polarization of the HERA electron beam comes naturally due to the Sokolov-Ternov effect, and is already used for the HERMES experiment. Polarizing the proton beam is a much more complicated task, and other sessions of this workshop were devoted to the machine aspects connected with this project. The physics prospects of a polarized HERA collider were investigated for the first time in a working group of the 1995/96 “Future Physics at HERA” workshop . The most important observables identified in this working group were the polarized structure function $`g_1(x,Q^2)`$, polarized weak structure functions, dijet production in polarized DIS and polarized photoproduction of jets. The working group established the measurability of all these observables, given an integrated luminosity of at least 200 pb<sup>-1</sup>. Encouraged by these results, a follow-on workshop “Physics with Polarized Protons at HERA” was organized in 1997 , where more elaborate studies of the measurability of different aspects of the proton spin structure at HERA were carried out. For present workshop, many of these analyses have been further refined, and a number of new processes, like deeply virtual compton scattering or spin asymmetries in diffraction and in leptoquark production, have been investigated in detail for the first time. A new aspect to spin physics at HERA is also the possibility of colliding the HERA proton beam with the polarized electron beam of a linear electron-positron collider, whose construction is currently considered at DESY. Several studies at this workshop have illustrated the physics prospects of this project. ## 2 Current and future experiments The HERMES collaboration is operating a polarized fixed target spectrometer with final state hadron identification in the HERA electron beam. In addition to measurements of individual polarized quark distributions from semi-inclusive asymmetries , this experiment has recently provided a first glimpse on the polarized gluon distribution by studying charged hadron production at high transverse momentum . In the near future, HERMES will study a variety of inclusive and semi-inclusive observables in polarized electron-nucleon scattering. In particular the semi-inclusive measurements at HERMES will yield new information on angular lepton-hadron correlations and on transverse momentum dependence (Collins effect) in fragmentation processes . Running at lower electron energy than HERMES, the polarized deep inelastic scattering programme at Jefferson Laboratory is aiming to study spin structure functions at large $`x`$ and/or low $`Q^2`$. This kinematic region is of particular interest for the determination of higher twist contributions to polarized structure functions, testing integral relations , lattice calculations and bounds on asymmetries . The COMPASS experiment , which is currently under construction at CERN, will start measuring inclusive and semi-inclusive observables in polarized lepton-nucleon scattering next year. With a lepton beam energy significantly above HERMES, the COMPASS experiment will be able to use charm production as a probe of the polarized gluon distribution. COMPASS will also carry out a broad programme of inclusive as well as semi-inclusive measurements on longitudinally and transversely polarized targets. A new domain for spin physics will open up with the commissioning of the Relativistic Heavy Ion Collider (RHIC) at BNL two years from now. RHIC can be operated with polarized proton beams, allowing to study polarized proton-proton collisions at $`\sqrt{s}=200\mathrm{}500`$ GeV with two multi-purpose collider detectors. The RHIC spin physics programme covers a wide variety of processes probing different aspects of the nucleon spin structure; first detailed simulations are now ongoing and have been reported to the workshop. Polarized quark distributions for individual flavours can be accessed at RHIC from massive gauge boson production, simulations indicate that this measurement should be relatively unproblematic . The polarized gluon distribution could be measured at RHIC from prompt photon or jet production, both channels are however not free from experimental as well as theoretical problems. In particular the direct photon channel has been subject of detailed simulation , demonstrating the experimental feasibility of measuring this observable. The theoretical interpretation of direct photon data is however problematic already in the unpolarized case, since the experimental isolation cuts, which are applied to define the photon, can hardly be matched in the theoretical calculation. The extraction of the gluon distribution from these data is therefore not free from ambiguities, direct photon data are therefore no longer used in unpolarized fits. The problem of photon isolation will be the same in polarized studies, thus casting doubt on the reliability of this measurement. Alternative photon isolation criteria could allow a better matching of experiment and theory. Using the isolated photon definition proposed in , a first case study for RHIC was carried out recently . Jet production at RHIC as a probe of the polarized gluon distribution has not yet been investigated in great detail, this process has the advantage of having large production rates, but suffers from the numerous competing partonic subprocesses. Data from HERMES, Jefferson Laboratory, COMPASS and RHIC will largely extend our knowledge on several aspects of the nucleon spin structure over the next few years. These measurements are however limited in their kinematical reach, and they will still leave many questions unanswered. All three polarized lepton-nucleon scattering experiments are working on fixed targets, and are therefore not able to access the behaviour of polarized structure functions at small $`x`$. Also, photoproduction at these experiments will largely be dominated by direct photon-nucleon interactions, thus not allowing studies of the spin structure of the photon. Measurements at the RHIC collider will also cover only the behaviour of the polarized parton distributions at large and medium $`x`$, which might in particular be problematic for an accurate determination of the first moment of the polarized gluon distribution. In unpolarized proton structure, similar aspects are of high interest, and could first be studied experimentally at the HERA electron-proton collider. ## 3 Prospects of Spin Physics at HERA The operation of HERA with polarized proton and electron beams would allow to study a wide variety of observables in polarized electron–proton collisions at $`\sqrt{s}=300`$ GeV. The physics prospects of this project were first studied in a working group of the 1995/96 workshop “Future Physics at HERA” and a follow-on workshop “Physics with Polarized Protons at HERA” in 1997. These workshops identified a number of key observables, which allow to probe several aspects of the proton and photon spin structure inaccessible at other experiments. These encouraging physics prospects also triggered much effort on machine studies for operating HERA with a polarized proton beam. At the present workshop, latest developments on machine aspects and physics prospects were shown. Concerning the physics studies, much effort has gone into optimization and refinement of the measurement of key observables as well as into studies of potential new channels. The most prominent new results are summarized in the following. ### 3.1 Spin Physics at small $`x`$ and large $`Q^2`$ The unpolarized electron-proton programme at HERA has extended the kinematical domain of deep inelastic scattering towards small values of the Bjorken scaling variable $`x`$ and large values of the invariant momentum transfer $`Q^2`$. Both limits probe the structure of the proton and the dynamics of QCD interactions at a previously unexplored level and give rise to a multitude of new phenomena, which have been studied in detail by the two HERA collider experiments over the past years. The kinematical reach of HERA towards large $`Q^2`$ allows perturbative QCD evolution of structure functions to be tested over a long range in $`Q^2`$, thus allowing for example for an indirect determination of the unpolarized gluon distribution from $`F_2(x,Q^2)`$. Likewise, operation of HERA with polarized protons would allow measurements of $`g_1(x,Q^2)`$ well beyond the kinematical reach of present fixed target experiments, thus providing important constraints to QCD fits and allowing for an indirect extraction of the polarized gluon distribution from QCD evolution of $`g_1(x,Q^2)`$. The prospects of this measurement have been investigated in detail already at previous workshops , some updates have now made. Concerning detector effects on the extraction of $`g_1(x,Q^2)`$ from the measured asymmetry, new studies of uncertainties induced by bin migration have been carried out , showing in particular that bin migration effects can be reduced by using the hadron method for the reconstruction of the event kinematics. To study the impact of HERA structure function data on a determination of polarized parton distributions from QCD fits, projected HERA data points have been included in global fits . A recent improvement to these studies is the inclusion of projected charged current data, which will allow for a more precise determination of the quark contribution to the proton spin . A new aspect to measurements of $`g_1(x,Q^2)`$ at HERA would be the possibility of colliding the HERA proton beam with the electron beam of a future linear collider, whose construction is currently considered at DESY. First studies on the kinematical reach of the option were presented . Already a linear collider with electron beam energy of 250 GeV would enlarge the kinematical reach of HERA by about an order of magnitude towards large $`Q^2`$ and small $`x`$, see Fig. 1. The behaviour of polarized structure functions at small $`x`$ is currently not known from experiment. Present $`g_1(x,Q^2)`$ data from the SMC collaboration reach down to $`x=0.0008`$, corresponding however to very small values of $`Q^2=0.2`$GeV<sup>2</sup>. The SMC data have motivated several attempts of a theoretical interpretation , which are however yet inconclusive. In the small-$`x`$ domain, one would ultimately expect the ln $`Q^2`$ resummation of QCD to break down, since terms proportional to ln $`x`$ should become of equal importance, thus requiring a reordering of the perturbative series. HERA measurements of the unpolarized structure function $`F_2(x,Q^2)`$ at small $`x`$ are however still consistent with the DGLAP evolution equations, based on ln $`Q^2`$ resummation. In this structure function, one finds the most singular terms at small $`x`$ to be of the form $`\alpha _s^n\mathrm{ln}^nx`$. In the polarized structure function $`g_1(x,Q^2)`$, even more singular terms of the form $`\alpha _s^n\mathrm{ln}^{2n}x`$ are present , resulting in a stronger enhancement at small $`x`$. At this workshop, a first attempt has been presented to combine small-$`x`$ resummation with the usual DGLAP evolution into a unified evolution equation. As a result, one observes a significant enhancement of the magnitude of $`g_1`$ at small $`x`$, compared to DGLAP evolution only. A similar enhancement is also expected in $`g_2(x,Q^2)`$ at small $`x`$ ; a measurement of this structure function seems however to be unlikely at HERA. About 10% of unpolarized deep inelastic scattering events at HERA yield a final state containing a diffractively scattered proton surrounded by a large rapidity gap. Since its first observation at HERA several years ago, diffraction in deep inelastic scattering has triggered a lot of theoretical effort towards its understanding. At present, it is still fair to say that an unambiguous interpretation of this phenomenon could not yet be achieved, since both perturbative as well as non-perturbative effects are expected to contribute with comparable magnitude. Theoretical studies on diffraction in polarized deep inelastic scattering were presented at this workshop. Estimations of non-perturbative contributions based on regge theory predict a diffractive asymmetry not exceeding $`10^4`$, while perturbative diffractive exchanges give rise to asymmetries of the order $`10^2`$, as can been seen in Fig. 2. The ratio of perturbative to non-perturbative contributions to diffraction appears therefore to be more favourable in the polarized than in the unpolarized case, in particular due to terms of the form $`\alpha _s^n\mathrm{ln}^{2n}x`$, which are present only in the polarized, but not in the unpolarized diffractive amplitude. Studies of particular final states in polarized diffractive scattering are also ongoing, large asymmetries are predicted for example in diffractive charm production. ### 3.2 Jet production in deep inelastic scattering The most promising direct probe of the polarized gluon distribution $`\mathrm{\Delta }g(x,Q^2)`$ at HERA is the measurement of asymmetries in dijet production in polarized deep inelastic scattering. Already the past workshops have proven that this observable allows for a determination of $`\mathrm{\Delta }g(x,Q^2)`$ down to $`x=0.003`$, almost an order of magnitude in $`x`$ below the reach of measurements at RHIC. Since current QCD fits predict an important contribution to the first moment of $`\mathrm{\Delta }g(x,Q^2)`$ from this region, this measurement appears to mandatory for a determination of the gluon contribution to the proton spin. Earlier analyses of dijet production in polarized deep inelastic scattering have now been extended considerably with the incorporation of next-to-leading order corrections to the parton level subprocesses . It could be shown that the gluon induced scattering process remains dominant source of dijet events also at NLO. Moreover, the non-trivial correlation between the proton energy fraction $`x_g`$ of the gluon contributing to the scattering process and the reconstructed $`x_{jets}`$ from the final state is now understood at NLO, thus allowing for a consistent extraction of the polarized gluon distribution at this order. The expected jet production asymmetry at next-to-leading order is displayed in Fig. 3. The impact of dijet data on QCD fits of polarized parton distributions has been investigated in detail . It could be shown that these data are in particular essential to constrain the shape of the gluon distribution. The kinematical reach of the dijet measurement would also be extended by almost another order of magnitude in $`x`$ for the option of colliding the electron beam from a linear collider with the HERA proton beam . As an alternative to the dijet measurement, it has been suggested to use current-target correlations to determine the polarized gluon distribution . This process has the advantage that no requirement on jets in the final state is made, thus including all inclusive deep inelastic scattering events in the analysis. The extraction of $`\mathrm{\Delta }g(x,Q^2)`$ from this process is however far less clean than from the dijet measurement, large systematic uncertainties must be expected. Jet production in the forward proton region has been suggested as a probe of QCD enhancement effects at small $`x`$ in the unpolarized case. Measurements of this observable turned out to be substantially above the expectations obtained by resumming only ln $`Q^2`$ terms, thus indicating the presence of large ln $`x`$ in this process. First studies of small-$`x`$ resummation effects in polarized forward jet production were presented at the workshop , indicating that small-$`x`$ resummation results in a characteristic increase of the cross section for this observable, which could therefore be used as a tool to study small-$`x`$ dynamics in polarized scattering. ### 3.3 Polarized Photoproduction Cross sections in electron-proton collisions become largest, if the virtuality of the photon mediating the interaction is small. In this photoproduction limit, one can approximate the electron-proton cross section as a product of a photon flux factor and an interaction cross section of a real photon with the proton. Many unpolarized photoproduction reactions are presently measured at HERA, and their study has continuously improved our knowledge on proton and photon structure as well as our understanding of the transition between real and virtual photons over the last years. Polarized photoproduction processes have already been in the focus of the past two workshops , where jet photoproduction was identified as the most promising channel for a determination of both the polarized gluon distribution in the proton and the polarized parton distributions in the photon. Jet production in the photon direction originates mainly from photon-gluon fusion processes, and thus reflects the gluon polarization in the proton. The situation is more involved in the proton direction, where most events are induced by the yet unknown resolved partonic content of the polarized photon. Given the polarized parton distributions in the proton to be known from other sources, jet photoproduction in the proton direction can be used to determine the polarized parton distributions in the resolved photon. Next-to-leading order QCD corrections to polarized photoproduction of jets have recently been calculated , resulting in larger perturbative stability of the predictions and enabling a consistent extraction of parton distributions at this order. Knowledge on parton distributions in the polarized photon is essential to exploit the full physics potential of electron-photon and photon-photon collisions at a future linear electron-positron collider, which is foreseen to operate with polarized beams. At present, it seems that polarized photoproduction of dijets at HERA is the unique process to determine these distributions. Given that jet production originates from several different partonic subprocesses, unfolding of all polarized quark and gluon distributions in the photon from dijet measurements is a very involved, if not impossible task. In the unpolarized case, one successfully applies the so-called effective parton approximation . In this approximation, one replaces the sum over different partonic subprocesses, each weighted with a different parton distribution, by one universal subprocess, weighted with an effective distribution. This effective distribution is a linear combination of parton distributions, where each distribution is weighted with the approximate importance of the corresponding subprocess. This approximation has now been extended to the polarized case , where it will help to quantify the expected precision of a polarized photon structure measurement at HERA. Polarized photoproduction of charmed quarks has been discarded as competitive probe of the polarized gluon distribution already at previous workshops due to the expected low charm tagging efficiency. The production of open charm and of bound states of charmed quarks has however received renewed interest at this workshop as testing ground for models for charm production and charm fragmentation . ### 3.4 Deeply virtual compton scattering The exclusive process $`\gamma ^{}p\gamma p`$ (deeply virtual compton scattering, DVCS) can be used to probe aspects of the proton structure beyond the observables usually studied in inclusive and semi-inclusive deep inelastic scattering . It particular, this reaction allows a measurement of non-forward (skewed) parton distributions, which are universal quantities expected to arise in diffractive and exclusive scattering processes. A direct measurement of the cross section for DVCS is however not possible at HERA, since this process is concealed by a large background from photon bremsstrahlung in ordinary electron-proton scattering. By measuring single spin asymmetries in exclusive photon production in the scattering of polarized electrons off unpolarized protons, it is however possible to access the interference term of the DVCS amplitude with the DIS amplitude. First numerical studies of this process have been presented at the workshop . These are indicating large single spin asymmetries, which will allow a determination of skewed parton distributions at HERA, once both collider experiments are operating with longitudinal electron polarization. Adding proton polarization would enable a measurement of polarized skewed parton distributions as well. ### 3.5 Searches for new phenomena If the upcoming high luminosity runs at HERA show some evidence for deviations from the Standard Model, polarization of the proton beam could help to determine the origin of this deviation. A case study on new physics at polarized HERA for leptoquark production has been reported at this workshop . Leptoquarks are objects carrying both lepton and quark quantum numbers, they can be created in lepton-quark scattering and would show us as peaks in the $`x`$-distribution of deep inelastic scattering events. According to their quantum numbers and their coupling structure to leptons and quarks, leptoquarks can be classified into different species . HERA could in principle detect leptoquarks with a mass up to almost its electron-proton centre-of-mass energy. This detection is already possible in unpolarized scattering. However, based on information from unpolarized scattering only, it will turn out to be very hard to determine the species of the observed leptoquark. In , it was demonstrated that polarization asymmetries in leptoquark production can be used extract the precise coupling structure of an observed leptoquark, and thus to determine its species. Fig. 4 illustrates the potential to discriminate different leptoquark species by measuring parity violating spin asymmetries at $`\sqrt{s}=380`$ GeV, which is about the maximum electron proton centre-of-mass energy that could be eventually reached at HERA. ### 3.6 Fixed target programme at HERA–$`\stackrel{}{\mathrm{N}}`$ In addition to the physics programme at the polarized $`ep`$ collider, it would be possible to study polarized proton-nucleon collisions at a fixed target experiment in the polarized HERA proton beam. This proposed experiment, HERA–$`\stackrel{}{\mathrm{N}}`$, would require a polarized internal nucleon target and a dedicated new spectrometer. It could add numerous hadron-hadron observables to the HERA spin physics programme. Very interesting results are expected here, which are fully complementary to the RHIC spin physics program. The two main issues are probes of the polarized parton distributions in double spin asymmetries, e.g. in Drell-Yan process or direct photon production and investigations into hadronic single spin asymmetries. At this workshop, several new investigations for single spin asymmetries at HERA–$`\stackrel{}{\mathrm{N}}`$ were presented: diffractive production of charmed baryons may be used to investigate the spin structure of diffractive interactions, instanton effects could show up in single spin asymmetries and pion production could serve as a probe of spin transfer in peripheral interactions . ## 4 Conclusions and Outlook A variety of new information on the spin structure of the nucleon can be expected over the next few years from the continuation of the HERMES experiment and from COMPASS and RHIC, which are currently under construction. These experiments will jointly provide the first data on the gluon polarization in the nucleon from several complementary observables, and supply new information on the quark contribution to the nucleon spin. Despite this wealth of new data to be expected, many questions in spin physics will still remain open. All experiments are limited in their kinematical reach, in particular for the determination of the gluon contribution to the proton spin. Moreover, none of these experiments is sensitive on the structure of the polarized photon - which is completely unknown at present, but of high importance for precision predictions of observables to be studied at future electron-photon or photon-photon colliders. Polarization of the HERA proton beam would allow the study of polarized electron-proton and photon-proton collisions at high energies, which would allow to measure several aspects of spin structure of proton and photon which are inaccessible elsewhere. Table 1 summarizes the information on polarized parton distributions in the nucleon that can be gained from present and future experiments. The first block summarizes the current status, the second block lists the new reactions that can be probed at ongoing and currently constructed experiments, while the last block illustrates the improvements that could be made at a polarized HERA collider. At this workshop, new results on the measurability of these key observables were presented: systematic uncertainties in the extraction of the structure function $`g_1(x,Q^2)`$ from asymmetry measurements are now better understood, next-to-leading order QCD corrections are included in the extraction of the gluon distribution from dijet production and the impact of including charged current HERA data in global fits has been demonstrated. Concerning the study of polarized photon structure at HERA, progress has been made with the introduction of an effective parton approximation, facilitating the extraction of parton distributions from measured asymmetries. The measurability of several new and potentially interesting observables has been demonstrated at this workshop: spin asymmetries in deep inelastic diffraction could probe the perturbative nature of the diffractive exchange, asymmetries in deeply virtual compton scattering might yield new insights into the proton structure and parity violating asymmetries could prove to be a valuable tool to study possible effects of new physics at HERA. A new aspect of spin physics at HERA is the possibility of colliding the polarized electron beam of a future linear collider with the polarized HERA proton beam. First studies on the physics prospects of this option were presented at the workshop, and work in this direction is ongoing. Given that the full potential of the linear collider itself can only be exploited if information on the polarized photon structure is available from other sources, this option highlights the complementarity of spin physics at HERA with other future projects at DESY.
no-problem/9908/astro-ph9908027.html
ar5iv
text
# Collisionless Halos Around Black Holes ## 1. Introduction In 1970, Victor Shvartsman introduced the idea of hot halos around slowly-accreting black holes. Today, we believe that these halos do exist around the supermassive black holes in the centers of non-AGN galaxies. The radio source Sgr A is the nearest example. The radiative efficiency of collisionless halos is determined by the electron heating mechanisms (§2). The rate of accretion is not necessarily given by the Bondi formula (§3). The plasma density and temperature profiles near the black hole are estimated in §4. We conclude in §5. ## 2. The Radiative Efficiency At low densities, the radiation is weak and the accreting gas is unable to cool. More precisely, the protons are unable to cool. The proton temperatures are close to $`0.1\mathrm{GeV}`$ near the last stable orbit. For realistic densities, the plasma is collisionless. Electrons synchrotron cool efficiently and are much colder than ions. Electrons radiate away nearly all the heat they receive. The radiative efficiency of accretion ($`\eta L/\dot{M}c^2`$, $`L`$ is the luminosity, $`\dot{M}`$ is the accretion rate) is determined by the electron heating processes. The electron heating processes depend on the magnetic field strength. The magnetic field should be close to equipartition with the gravitational energy. Shvartsman (1971) explains this as follows. For a nearly spherical accretion, assuming perfect flux-freezing, the radial magnetic field is $`r^2`$, where $`r`$ is the radius. The magnetic energy density is $`ϵ_mr^4`$. Assuming a free fall, the velocity is $`r^{1/2}`$. From continuity, the density is $`r^{3/2}`$. The gravitational energy density is $`ϵ_gr^{5/2}ϵ_m`$. But the magnetic energy cannot exceed the gravitational energy, therefore $`ϵ_mϵ_g`$. When magnetic fields are close to equipartition, electrons should receive approximately as much heat as ions, that is about $`0.1\dot{M}c^2`$. There are two different mechanisms of electron heating. First, the viscous heat is evenly distributed between electrons and ions (Gruzinov 1988, Quataert & Gruzinov 1999). Second, reconnection should be an efficient electron heater (Blandford 1998, Quataert & Gruzinov 1999). If the electron heating is not negligible, as these papers suggest, the radiative efficiencies as small as $`10^5`$ are impossible. Then the observed low luminosity of the non-AGN supermassive black holes should be explained by the small rate of accretion. Blandford & Begelman (1998) believe that the small accretion rate is explained by the wind loses. A different explanation is proposed in §3. ## 3. The Accretion Rate The rate of spherical accretion onto an object of mass $`M`$ is given by the Bondi formula $$\dot{M}_{\mathrm{Bondi}}\rho c_sR_A^2,$$ (1) where $`\rho `$ and $`c_s`$ are the density and the speed of sound of the accreting gas at a large distance from the object, and $`R_A=2GM/c_s^2`$ is the accretion radius. For a supersonic wind, $`c_s`$ is replaced by the wind velocity. The Bondi formula is widely used in astronomy. This formula was derived assuming that the flow is: (i) spherical, (ii) inviscid, (iii) adiabatic. Although neither of these assumptions is realistic, the Bondi formula is a good order of magnitude estimate for the rate of accretion in many astronomically relevant cases (§3.1). We show, however, that in the case of low-density accretion onto a black hole, the Bondi formula can overestimate the rate of accretion by orders of magnitude (§3.2, §3.3) <sup>1</sup><sup>1</sup>1Most of the accretion disk papers treat $`\dot{M}`$ as an adjustable parameter. Here we actually calculate it, just like Bondi did, but under different assumptions.. For low-density quasi-spherical accretion onto a black hole, the accreting matter is a hot collisionless plasma. This plasma is an ideal heat conductor along the magnetic field lines. Heat conduction strongly violates the adiabaticity assumption used to derive the Bondi formula. In the absence of heat conduction, thermal energy of the plasma is lost through the event horizon. When heat conduction becomes important, some of this heat is transfered out, away from the hole. The temperature and pressure of the accretion flow increase over what have been the case for the adiabatic solution. The increased pressure opposes the gravitational attraction of the hole and impedes the accretion. In §3.1 we derive the Bondi formula and explain why it works for real (non-spherical, viscous, non-adiabatic, but relatively cool) flows. In §3.2 we show that if the effective adiabatic index of the flow $`\gamma _{\mathrm{eff}}d\mathrm{ln}p/d\mathrm{ln}\rho `$ is greater than 5/3, the accretion slows down. We then show that heat conduction leads to $`\gamma _{\mathrm{eff}}>5/3`$, and derive a formula for the accretion rate in the presence of heat conduction. In §3.3 we present a more rigorous derivation of the same formula. When heat conduction is important, the rate of quasi-spherical accretion onto a black hole is $$\dot{M}\dot{M}_{\mathrm{Bondi}}\left(\frac{R_S}{R_A}\right)^\alpha ,$$ (2) where $`R_S=2GM/c^2`$ is the Schwarzschild radius <sup>2</sup><sup>2</sup>2Relativistic effects are much smaller than the effects we are considering in this paper. We treat our black holes as Newtonian holes in space, and neglect the special-relativistic effects., and $`\alpha 1`$ is a model-dependent dimensionless number. To illustrate the astrophysical applications of the accretion rate formula (2), consider the case of Sgr A, which is believed to be a $`2.5\times 10^6M_{}`$ black hole located at the center of our Galaxy (Genzel et al 1994, Ghez et al 1998). The accretion luminosity of Sgr A is believed to be $`L10^{37}\mathrm{erg}/\mathrm{s}`$. When stellar winds in the vicinity of the galactic center collide and shock, they produce a gas of $`c_s1000\mathrm{k}\mathrm{m}/\mathrm{s}`$ and $`\rho 10^{20}\mathrm{g}/\mathrm{cm}^3`$ (Coker & Melia 1997). The Bondi accretion rate of this gas is $`\dot{M}_{\mathrm{Bondi}}10^{21}\mathrm{g}/\mathrm{s}`$. Thus $`\dot{M}_{\mathrm{Bondi}}c^210^{42}\mathrm{erg}/\mathrm{s}`$ is five orders of magnitude higher than the actual luminosity. We (Gruzinov 1998, Quataert & Gruzinov 1998) and others (Shvartsman 1971, Meszaros 1975, Blandford 1998) have argued that such low radiative efficiencies are unrealistic. It is then natural to assume that Sgr A accretes at a much smaller rate (Blandford & Begelman 1998). If we use the estimate (2), and, for illustrative purposes, assume $`\alpha =0.4`$, the estimated accretion rate is reduced by a factor of 100. The radiative efficiency is then $`0.1\%`$. This radiative efficiency is still small, but it might be reasonable because electrons might be heated less than ions, and even hot electrons might radiate inefficiently when the plasma is rarefied (Quataert & Narayan 1998). ### 3.1. The rate of spherical accretion according to Bondi The Bondi formula can be explained as follows. At a distance $`rR_A`$ from the massive object, the gravitational energy of the gas particles is comparable to the thermal energy. It is plausible that all these particles fall onto the black hole. Further assume that at $`rR_A`$ the radial velocity is $`c_s`$ and the density is $`\rho `$. We then get the mass accretion rate given by (1). Is the pressure build-up on the way into the massive object important? Can the pressure forces counter the gravitational attraction? We can answer this question by first neglecting the pressure forces at small radii, $`R_A`$. Then the radial velocity is equal to the free fall velocity, $`r^{1/2}`$. From mass continuity, the density is $`r^{3/2}`$. The pressure is $`r^{3\gamma /2}`$, where $`\gamma `$ is the adiabatic index of the gas. The acceleration from the pressure gradient is $`r^{(3\gamma 1)/2}`$. The gravitational acceleration is $`r^2`$. So long as $`\gamma <5/3`$, the pressure is not able to resist gravity. This qualitative derivation explains why the Bondi formula should work well if the accreting gas has an effective adiabatic index $`\gamma _{\mathrm{eff}}5/3`$. In this case, the pressure forces cannot impede accretion, and the gas from $`rR_A`$ must accrete at about the Bondi rate (1). Numerical simulations of a non-spherical accretion confirm the Bondi formula (Coker & Melia 1997). ### 3.2. $`\gamma _{\mathrm{eff}}>5/3`$ Suppose that the accreting matter is a collisionless plasma. Suppose that radiation is negligible. We will show that heat conduction increases the effective adiabatic index to $`\gamma _{\mathrm{eff}}>5/3`$. Qualitatively, the heat conduction reduces the loss of thermal energy through the event horizon. The temperature of the flow increases faster than for an adiabatic plasma, which means that $`\gamma _{\mathrm{eff}}>5/3`$. Quantitatively, $`TdS=dQ`$, or $$\rho Tv_rs=r^2_r(r^2\kappa _rT).$$ (3) Here $`\rho `$ is the density, $`T`$ is the temperature, $`v`$ is the inflow (positive) velocity, $`s`$ is the entropy per unit mass, $`\kappa `$ is the thermal conductivity. We assume that the thermal conductivity $`\kappa `$ is given by a “Shakura-Sunayev-like” formula $$\kappa =m^1\alpha \rho rv,$$ (4) where $`\alpha 1`$ is a dimensionless positive constant, $`m`$ is the molecular mass. The Shakura-Sunayev-like prescription is plausible because the characteristic size and frequency of the turbulent structures should be $`r`$ and $`v/r`$. The thermal conductivity might be suppressed for a supersonic inflow, but our accretion flow is subsonic. With $`s=m^1\mathrm{ln}(T^{3/2}/\rho )`$, equation (3) gives $$\rho Tv\left(\frac{3}{2}\frac{T^{}}{T}\frac{\rho ^{}}{\rho }\right)+\alpha r^2(r^3\rho vT^{})^{}=0,$$ (5) where the prime denotes the r-derivative. For a stationary flow $`\rho vr^2=\mathrm{const}`$, and $$T\left(\frac{3}{2}\frac{T^{}}{T}\frac{\rho ^{}}{\rho }\right)+\alpha (rT^{})^{}=0.$$ (6) If $`\alpha `$ is small, we can use the adiabatic relationship in the last term, $`rT^{}=T`$. Then equation (6) gives $$T^{\frac{3}{2}\alpha }\rho ,$$ (7) that is $$\gamma _{\mathrm{eff}}=\frac{52\alpha }{32\alpha }.$$ (8) Since $`\gamma _{\mathrm{eff}}>5/3`$, pressure will impede the accretion. We now estimate the suppressed accretion rate. Assume that at the event horizon the temperature is close to $`mc^2`$. The density at the hole $`\rho _h`$ is given by the adiabatic law $$\frac{\rho _h}{\rho }\left(\frac{c^2}{c_s^2}\right)^{\frac{1}{\gamma _{\mathrm{eff}}1}}\left(\frac{R_A}{R_S}\right)^{\frac{1}{\gamma _{\mathrm{eff}}1}}.$$ (9) The rate of accretion is $`\dot{M}R_S^2c\rho _h`$, and from (8) and (9) we get (2). ### 3.3. A solvable model In this section we rigorously derive the accretion rate when heat conduction is important, equation (2). We assume a spherically symmetrical gas inflow. The radiative cooling is neglected. The flow is taken to be inviscid but non-adiabatic, i.e. we include thermal conduction and neglect viscosities. In terms of suppressing the rate of accretion, heat conduction is more important than non-sphericity or viscosity. Non-spherical accretion is known to proceed at about the Bondi rate (Coker & Melia 1997). Viscosity suppresses the accretion rate, but only weakly, because when the inflow velocity decreases, the effects of viscosity decrease faster than the adiabatic effects (see also Turolla & Nobili 1989). We treat the flow using a fluid approximation, but we should remember that in reality we are dealing with magnetized collisionless plasmas. However, the fluid approximation obeys the correct set of conservation laws, and should be close to reality. The qualitative conclusions of our analysis should be taken seriously. Consider a non-relativistic radial inflow of the $`\gamma =5/3`$ gas. The gas density $`\rho `$, the inflow (positive) velocity $`v`$, and temperature $`T`$ depend on the radial coordinate $`r`$ only, and satisfy the stationary equations of continuity, Euler’s, and the thermal conduction. $$\dot{M}=4\pi r^2\rho v,$$ (10) $$vv^{}=\frac{(\rho T)^{}}{m\rho }\frac{GM}{r^2},$$ (11) $$\frac{\rho Tv}{m}\left(\frac{\rho ^{}}{\rho }\frac{3}{2}\frac{T^{}}{T}\right)=r^2(\kappa r^2T^{})^{}.$$ (12) We will use dimensionless units with $`GM=1`$, and the molecular mass $`m=1`$. The boundary condition at the surface of the compact object is unknown, but it is irrelevant. For concreteness, assume $`T+\beta rT^{}=0`$ at $`r=R_S`$, where $`\beta `$ is a dimensionless constant. At $`r=\mathrm{}`$, we assume $`T=1`$, so that $`R_A1`$. The boundary condition for $`\rho `$ is also irrelevant, because the system of equations is invariant under $`\rho \lambda \rho `$, $`\dot{M}\lambda \dot{M}`$. Our aim is to find the maximal possible value of the dimensionless mass accretion rate $`\dot{m}\dot{M}/\rho (\mathrm{})`$, for which a smooth solution of the system exists. The ratio $`\dot{M}/\dot{M}_{\mathrm{Bondi}}`$ is equal to this maximal value. Obviously $`\dot{m}1`$ if $`R_S1`$. We need to find out how $`\dot{m}`$ depends on $`R_S`$ when $`R_S1`$. Scale out $`\dot{M}`$ by $`\rho \dot{M}\rho /4\pi `$, insert (4) and (10) into (12): $$\rho vr^2=1,$$ (13) $$vv^{}=\rho ^1(\rho T)^{}r^2,$$ (14) $$T\left(\frac{\rho ^{}}{\rho }\frac{3}{2}\frac{T^{}}{T}\right)=\alpha (rT^{})^{}.$$ (15) Using (15), integrate (14) and obtain a system of two first-order differential equations $$\frac{1}{2}\frac{1}{r^4\rho ^2}+\frac{5}{2}T+\alpha rT^{}=\frac{1}{r}+\frac{5}{2},$$ (16) $$\frac{\rho ^{}}{\rho }\left(T\frac{1}{r^4\rho ^2}\right)=\frac{1}{r^2}+\frac{2}{r^5\rho ^2}T^{}.$$ (17) We need to find the dependence on $`R_S1`$ of the maximal possible value of $`\dot{m}1/\rho (\mathrm{})`$ for which a smooth solution of the system exists. For small $`r`$, we can neglect $`5/2`$ in the right hand side of (16). Then, denoting $`\tau =rT`$, $`f=(r^3\rho ^2)^1`$, $`x=\mathrm{log}r`$, we obtain $$2\alpha \frac{d\tau }{dx}=2(52\alpha )\tau f,$$ (18) $$\frac{df}{dx}=f\frac{2+2d\tau /dx5\tau f}{\tau f}.$$ (19) This system has a stable equilibrium point $`\tau =2(52\alpha )^1`$, $`f=0`$ (it is assumed that $`\alpha <5/2`$). In the vicinity of this point, $$\frac{df}{dx}=2\alpha f.$$ (20) The system (18), (19) is an accurate approximation only for $`r1`$, i. e., for an x-duration $`\delta x\mathrm{log}(1/R_S)`$. The quantity $`f`$ decreases according to (20) for an x-duration $`\delta x=\mathrm{log}(1/R_S)\mathrm{const}`$. Therefore the final density is proportional to $`(R_S)^\alpha `$. Since $`\dot{m}1/\rho (\mathrm{})`$, the scaling law (2) is proven. To check the answer, we integrated equations (16), (17) numerically. ## 4. The Density and Temperature of the Collisionless Halo The standard Bondi solution and the new solution of §3 predict different densities near the black hole. In both cases the temperature is close to virial $$T(r)m_pc^2\frac{R_S}{r}$$ (21) The density is given by the adiabatic. For the Bondi solution $$n(r)n_{\mathrm{}}\left(\frac{R_A}{r}\right)^{3/2},$$ (22) where $`n_{\mathrm{}}`$ is the density at a large distance from the hole (at accretion radius). When heat conduction is important, we use the effective value of the adiabatic index. This gives $$n(r)n_{\mathrm{}}\left(\frac{R_A}{r}\right)^{3/2\alpha }.$$ (23) For Sgr A, under the assumption of §3, the new density is two orders of magnitude smaller. The density profile should, in principle, be observable. The deviations from the $`r^{3/2}`$ law will prove that collsionless halos around black holes are not ideal. ## 5. Discussion The low luminosity of the supermassive black holes existing in the centers of most galaxies (Richstone 1998) is traditionally explained by the extremely low radiative efficiency of the spherical accretion. We offer an alternative explanation: these black holes accrete at a rate much smaller than the Bondi rate. We have explained that electron heating and the radiative efficiency of the collisionless halos around black holes should be non-negligible. Then the observed low luminosity of many supermassive black holes should be due to the small accretion rate. We generalized the Bondi formula by including thermal conductivity. Thermal conductivity heats the flow, by transporting heat from the event horizon outward into the plasma. As the result of heat conduction, the pressure rises and impedes the accretion flow. #### Acknowledgments. This work was supported by NSF PHY-9513835. I thank Eliot Quataert, John Bahcall, and Martin Rees for useful discussions. ## References Blandford, R. D., 1998, in ”Accretion Processes in Astrophysical Systems: Some Like It Hot”, eds. S. Holt and T. Kallman Blandford, R. D., & Begelman, M. C., 1998, MNRAS, submitted (astro-ph/9809083) Coker, R. F., & Melia, F., 1997, ApJ, 488, L149 Genzel, R., Hollenbach, D., & Townes, C. H., 1994, Rep. Prog. Phys., 57, 417 Ghez, A. M., Klein, B. L., Morris, M., & Becklin, E. E., 1998, ApJ, 509, 678 Gruzinov, A., 1998, ApJ, 501, 787 Meszaros, P., 1975, A& A, 44,59 Quataert, E., & Gruzinov, A., 1999, ApJ, 520, 248 Quataert, E., & Narayan, R., 1998, ApJ, submitted (astro-ph/9810136) Richstone, D., 1998, in Proceedings of the Second International LISA Symposium on Gravitational Waves, ed. W. Folkner (astro-ph/9810379) Shvartsman, V. F., 1971, Soviet Astronomy - AJ, 15, 377 Turolla, R., & Nobili, L., 1989, ApJ, 342, 982
no-problem/9908/hep-ph9908269.html
ar5iv
text
# References Is there a unique thermal source of dileptons in Pb(158 A$``$GeV) + Au, Pb reactions? K. Gallmeister<sup>a</sup>, B. Kämpfer<sup>a</sup>, O.P. Pavlenko<sup>a,b</sup> <sup>a</sup>Forschungszentrum Rossendorf, PF 510119, 01314 Dresden, Germany <sup>b</sup>Institute for Theoretical Physics, 252143 Kiev - 143, Ukraine Abstract An analysis of the dilepton measurements in the reactions Pb(158 A$``$GeV) + Au, Pb points to a unique thermal source contributing to the invariant mass and transverse momentum spectra. Effects of the flow pattern are discussed. Introduction: Dileptons are penetrating probes which carry nearly undisturbed information about early, hot and dense matter stages in relativistic heavy-ion collisions. Some effort, however, is needed for disentangling the various sources contributing to the observed yields and for identifying the messengers from primordial states of strongly interacting matter. The dielectron spectra for the reaction Pb(158 A$``$GeV) + Au measured by the CERES collaboration cannot be described by a superposition of $`e^+e^{}`$ decay channels of final hadrons, i.e. the hadronic cocktail. A significant additional source of dielectrons must be there. Since the data cover mainly the invariant mass range $`M<`$ 1.5 GeV the downward extrapolation of the Drell-Yan process is not an appropriate explanation. Also correlated semileptonic decays of open charm mesons have been excluded . As a widely accepted explanation, a thermal source is found to account for the data (cf. and further references therein, in particular with respect to in-medium effects and chiral symmetry restoration). Very similar, the NA50 collaboration has found, for the reaction Pb(158 A$``$GeV) + Pb, that the superposition of Drell-Yan dimuons and open charm decays does not explain the data in the invariant mass range 1.5 GeV $`<M<`$ 2.5 GeV . Final state interactions or abnormal charm enhancement have been proposed as possible explanations. Here we try to explain the NA50 measurements by a more apparent idea , namely a thermal source. We present a schematic model which at the same time describes the CERES and NA50 data. The model: Since we include the corresponding detector acceptances a good starting point for Monte Carlo simulations is the differential dilepton spectrum $$\frac{dN}{p_{\mathrm{\hspace{0.17em}1}}dp_{\mathrm{\hspace{0.17em}1}}p_{\mathrm{\hspace{0.17em}2}}dp_{\mathrm{\hspace{0.17em}2}}dy_1dy_2d\varphi _1d\varphi _2}=d^4Qd^4x\frac{dR}{d^4Qd^4x}\delta ^{(4)}(Qp_1p_2),$$ (1) where $`Q=p_1+p_2`$ is the pair four-momentum, $`p_{1,2}`$ are the individual lepton four-momenta composed of transverse momenta $`p_{\mathrm{\hspace{0.17em}1},2}`$ and rapidities $`y_{1,2}`$ and azimuthal angles $`\varphi _{1,2}`$. Here we extensively employ the quark - hadron duality and base the rate $`R`$ on the lowest-order quark - antiquark ($`q\overline{q}`$) annihilation rate (cf. ) $$\frac{dR}{d^4Qd^4x}=\frac{5\alpha ^2}{36\pi ^4}\mathrm{exp}\left\{\frac{uQ}{T}\right\},$$ (2) where $`u(x)`$ is the four-velocity of the medium depending on the space-time as also the temperature $`T(x)`$ does. Note that, due to Lorentz invariance, $`u`$ necessarily enters this expression. The above rate is in Boltzmann approximation, and a term related to the chemical potential is suppressed. As shown in the $`q\overline{q}`$ rate deviates from the hadronic one at $`M<`$ 300 MeV, but in this range the Dalitz decays dominate anyhow; in addition, in this range the thermal yield is strongly suppressed by the CERES acceptance. In the kinematical regions we consider below, the lepton masses can be neglected. Performing the space-time and momentum integrations in eqs. (1, 2) one gets $$\frac{dN}{dp_{\mathrm{\hspace{0.17em}1}}dp_{\mathrm{\hspace{0.17em}2}}dy_1dy_2d\varphi _1d\varphi _2}=\frac{5\alpha ^2}{72\pi ^5}p_{\mathrm{\hspace{0.17em}1}}p_{\mathrm{\hspace{0.17em}2}}_{t_i}^{t_f}𝑑tV(t)E,$$ (3) $`E`$ $`=`$ $`\{\begin{array}{cc}\mathrm{exp}\left\{\frac{M_{}\mathrm{cosh}Y\mathrm{cosh}\rho (r,t)}{T(r,t)}\right\}\frac{\mathrm{sinh}\xi }{\xi }\hfill & \text{for}3D,\hfill \\ K_0\left(\frac{M_{}\mathrm{cosh}\rho (r,t)}{T(r,t)}\right)I_0\left(\frac{Q_{}\mathrm{sinh}\rho (r,t)}{T(r,t)}\right)\hfill & \text{for}2D,\hfill \end{array}`$ (6) $`V(t)`$ $`=`$ $`\{\begin{array}{cc}4\pi 𝑑rr^2\hfill & \text{for}3D,\hfill \\ t𝑑rr\hfill & \text{for}2D,\hfill \end{array}`$ (9) where $`V(t)`$ acts on $`E`$, and $`3D`$ means spherical symmetric expansion, while $`2D`$ denotes the case of longitudinal boost-invariant and cylinder-symmetrical transverse expansion; the quantity $`\xi `$ is defined as $`\xi =T^1\mathrm{sinh}\rho \sqrt{M_{}^2\mathrm{cosh}Y^2M^2}`$, and $`\rho (r,t)`$ is the radial or transverse expansion rapidity; $`K_0`$ and $`I_0`$ are Bessel functions . The components of the lepton pair four-momentum $`Q=(M_{}\mathrm{cosh}Y,M_{}\mathrm{sinh}Y,\stackrel{}{Q}_{})`$ are related to the individual lepton momenta via $`M_{}^2`$ $`=`$ $`p_{\mathrm{\hspace{0.17em}1}}^2+p_{\mathrm{\hspace{0.17em}2}}^2+2p_{\mathrm{\hspace{0.17em}1}}p_{\mathrm{\hspace{0.17em}2}}\mathrm{cosh}(y_1y_2),`$ (10) $`\stackrel{}{Q}_{}`$ $`=`$ $`\stackrel{}{p}_{\mathrm{\hspace{0.17em}1}}+\stackrel{}{p}_{\mathrm{\hspace{0.17em}2}},`$ (11) $`M^2`$ $`=`$ $`M_{}^2Q_{}^2,`$ (12) $`\mathrm{tanh}Y`$ $`=`$ $`{\displaystyle \frac{p_1\mathrm{sinh}y_1+p_2\mathrm{sinh}y_2}{p_1\mathrm{cosh}y_1+p_2\mathrm{cosh}y_2}}.`$ (13) It turns out that the shape of the invariant mass spectrum $`dN/(dMdY|_{|Y|<0.5}dtdV(t))`$, which is determined only by the emissivity function $`E`$, does not depend on the flow rapidity $`\rho `$ in the 2D case , and in the 3D case for $`T`$ = 120 $`\mathrm{}`$ 220 MeV and $`\rho <0.6`$ there is also no effect of the flow. The analysis of transverse momentum spectra of various hadrons species point to an average transverse expansion velocity $`\overline{v}_{}`$ 0.43 at kinetic freeze-out, while a combined analysis of hadron spectra including HBT data yields $`\overline{v}_{}`$ 0.55 . Therefore, $`\rho <0.6`$ is the relevant range for the considered reactions. We note further that the invariant mass spectra $`dN/(dMdY|_{|Y|<0.5}dtdV(t))`$ for the 3D and 2D cases differ only marginally. Relying on these findings one can approximate the emissivity function $`E`$ by that of a ”static” source at midrapidity, as appropriate only for symmetric systems, $$E=\mathrm{exp}\left\{\frac{M_{}\mathrm{cosh}Y}{T(t)}\right\},$$ (14) thus getting rid of the peculiarities of the flow pattern. In contrast to the invariant mass spectra, the transverse momentum or transverse mass spectra are sensitive to the flow pattern , in general. A value of $`\rho =`$ 0.4, for example, causes already a sizeable change of the shape of the spectra $`dN/(dQ_{}dY|_{|y|<0.5}dtdV(t))`$ compared to $`\rho =0`$, in particular in the large-$`Q_{}`$ region. The differences between the 2D and 3D cases are not larger than a factor of 2 and, in a restricted $`Q_{}`$ interval, can be absorbed in a renormalization. The most striking difference of the 2D and 3D scenarios is seen in the rapidity spectrum: for 2D it is flat, while in the 3D case it is localized at midrapidity (values of $`\rho <`$ 0.6 also do not change the latter rapidity distribution). Below we shall discuss which space is left to extract from the dilepton data in restricted acceptance regions hints for the flow pattern when the other dilepton sources are also taken into account. Comparison with data: In line with the above arguments we base our rate calculations on eqs. (3, 14) and use the parameterizations $`T`$ $`=`$ $`(T_iT_{\mathrm{}})\mathrm{exp}\left\{{\displaystyle \frac{t}{t_2}}\right\}+T_{\mathrm{}},`$ (15) $`V`$ $`=`$ $`N\mathrm{exp}\left\{{\displaystyle \frac{t}{t_1}}\right\}.`$ (16) with $`T_i=`$ 210 MeV, $`T_{\mathrm{}}=`$ 110 MeV, $`t_1=`$ 5 fm/c, $`t_2=`$ 8 fm/c, $`N=\frac{A+B}{2.5n_0}`$ with $`A,B`$ as mass numbers of the colliding systems and $`n_0=`$ 0.17 fm<sup>-3</sup>. We stop the time evolution at $`T_f=`$ 130 MeV. In fig. 1 we show the comparison with the preliminary CERES data applying the appropriate acceptance . One observes a satisfactory overall agreement of the sum of the hadronic cocktail and the thermal contribution with the data. It is the thermal contribution which fills the hole of the cocktail around $`M=500`$ MeV in the invariant mass distribution in fig. 1a. In the mass bin $`M=`$ 0.25 $`\mathrm{}`$ 0.68 GeV the thermal yield is seen (fig. 1b) to dominate at small values of the transverse momentum $`Q_{}`$. In this region of $`Q_{}`$ transverse flow effects are not important. The large-$`Q_{}`$ spectrum is dominated by the cocktail. For higher-mass bins the thermal yield in the region of the first two data points is nearly as strong as the cocktail and rapidly falls then at larger values of $`Q_{}`$ below the cocktail. Therefore, the flow effects turn out to be of minor importance for the present analysis, since within our framework the transverse flow shows up at larger values of $`Q_{}`$. The question now is whether the same thermal source model accounts also for the NA50 data . The available NA50 data are not yet efficiency corrected and the efficiency matrix is not accessible. To have a reference we proceed as follows. According to the detailed analysis of the most central collisions, the shapes of the $`M`$ and $`Q_{}`$ spectra for invariant masses below the $`J/\psi `$ peak are described by the Drell-Yan yield $`+`$ 3 $`\times `$ the yield from correlated semileptonic decays of open charm mesons, both ones generated with PYTHIA . We have checked that our PYTHIA simulations coincide with results of the NA50 collaboration when applying the acceptance cuts. Our $`K`$ factors are adjusted by a comparison with Drell-Yan data and identified open charm data (cf. for the data compilation). Some confidence in our procedure is gained by correctly reproducing the NA50 interpretation of the NA38 data , which are efficiency corrected and which we can directly analyze: the NA38 data are described by the Drell-Yan yield $`+`$ 1.45 $`\times `$ the open charm contribution; the absolute normalization is obtained from a fit of the Drell-Yan region beyond the $`J/\psi `$. To get the yield for Pb + Pb collisions from PYTHIA the overlap function $`T_{AA}=`$ 31 mb<sup>-1</sup> is used. The resulting spectra (within the kinematical cuts for the NA50 acceptance) are displayed in figs. 2a and 2b by open squares. One observes that both the continuum part of the invariant mass spectrum (without the $`J/\psi `$ contribution) and the transverse momentum spectrum for the mass bin $`M=`$ 1.5 $`\mathrm{}`$ 2.5 GeV are nicely reproduced by the sum of Drell-Yan, open charm and thermal contributions. The thermal yield dominates again at not too large values of $`Q_{}`$ where transverse flow effects can be neglected. Therefore, it seems that from present dilepton measurements the transverse flow can hardly be inferred. Summary and discussion: In summary we have shown that a very simplified, schematic model for thermal dilepton emission, with parameters adjusted to the CERES data, also accounts for the measurements of the NA50 collaboration. A direct comparison with the NA50 data is not possible as long as the efficiency matrix is not at disposal. Nevertheless, our study points to a common thermal source seen in different phase space regions in the dielectron and dimuon channels. This unifying interpretation of different measurements has to be contrasted with other proposals of explaining the dimuon excess in the invariant mass region 1.5 $`\mathrm{}`$ 2.5 GeV either by final state hadronic interactions or an abnormally large open charm production. The latter one, in principle, can be checked experimentally , thus improving our understanding of dilepton sources. Due to the convolution of the local matter emissivity and the space-time history of the whole matter and the general dependence on the flow pattern, it is difficult to decide which type of matter (deconfined or hadron matter) emits really the dileptons. Our model is not aimed at answering this question. Instead, with respect to chiral symmetry restoration, we apply the quark - hadron duality as a convenient way to roughly describe the dilepton emissivity of matter by a $`q\overline{q}`$ rate, being aware that higher-order QCD processes change this rate (to some extent this might be included in a changed normalization $`N`$). In further investigations a more reliable rate together with a more detailed space-time evolution must be attempted and chemical potentials controlling the baryon and pion densities must be included. In this line the model has to be improved before using it for analyzing the dilepton yields in the asymmetric reactions S(200 A$``$GeV) + X measured by the NA38, HELIOS-3, and CERES collaborations, where also details of the rapidity distribution become important. Acknowledgments: Stimulating discussions with P. Braun-Munzinger, W. Cassing, O. Drapier, Z. Lin, U. Mosel, E. Scomparin, J. Rafelski, J. Stachel, and G. Zinovjev are gratefully acknowledged. O.P.P. thanks for the warm hospitality of the nuclear theory group in the Research Center Rossendorf. The work is supported by BMBF grant 06DR829/1.
no-problem/9908/cond-mat9908074.html
ar5iv
text
# Impact of magnetic frustration on the Mott transition within a slave-boson mean-field theory \[ ## Abstract We investigate the paramagnetic-metal-to-antiferromagnetic-metal and antiferromagnetic-metal-to-antiferromagnetic-insulator transitions using a slave-boson mean-field theory. To this effect, we discuss the ground state of the half-filled Hubbard model as a function of $`t^{}/t`$ and correlation strength $`U`$, where $`t`$ and $`t^{}`$ are the hopping amplitudes between nearest and next-nearest neighbors, respectively. The metal-insulator transition at a critical $`U_{\text{MIT}}`$ is of second order for small levels of magnetic frustration, $`t^{}/t<0.06`$, and of first order for large ones, $`t^{}/t>0.06`$. The insulator is always antiferromagnetically ordered, while the metal exhibits a second-order transition from a paramagnetic to an antiferromagnetic state up to $`t^{}/t=0.14`$, as $`U`$ is increased. We also contrast these findings with what we obtain in Hartree-Fock approximation. \] The correlation-driven metal-insulator transition, or Mott transition , observed in materials such as V<sub>2</sub>O<sub>3</sub> and NiS<sub>2-x</sub>Se<sub>x</sub> , is a nonperturbative problem usually tackled within the Hubbard model of strongly correlated electrons. This model describes itinerant electrons subject to an on-site repulsion $`U`$ comparable or greater than the bare bandwidth $`2D`$. In an early work, Brinkman and Rice investigated the Mott transition from the metallic side using Gutzwiller’s variational scheme . In this approximation, the metal is described as a strongly renormalized Fermi liquid. A low-energy scale $`ZD`$ ($`Z`$ is the quasiparticle residue) collapses linearly in $`U`$ as the Mott transition, occuring at a critical $`U_{\text{BR}}`$, is approached from the metallic side. $`ZD`$ is a measure for the renormalized Fermi energy. We investigate the implications of antiferromagnetic long-range order on either side of the Mott transition. To this effect, we introduce magnetic frustration, which helps stabilize an antiferromagnetic metallic phase for not too large levels of frustration and causes the insulating side to favor antiferromagnetic long-range order . We determine the phase diagram and the orders of the transitions in mean-field theory using a slave-boson technique. Our work complements previous studies and adds some analytical insights on the interplay of electron-electron correlations and magnetism. The single-band Hubbard model is given by $$\widehat{H}=\underset{ij\sigma }{}t_{ij}c_{i\sigma }^+c_{j\sigma }+U\underset{i}{}\widehat{n}_i\widehat{n}_i,$$ (1) where we take the amplitudes $`t_{ij}`$ to be nonzero only between nearest and next-nearest neighbors, in which cases they equal $`t`$ and $`t^{}`$, respectively. $`c_{i\sigma }^+`$ and $`c_{i\sigma }`$ are creation and annihilation operators for an electron of spin $`\sigma `$ at site $`i`$, and $`\widehat{n}_{i\sigma }c_{i\sigma }^+c_{i\sigma }`$. In this work, we consider the two-dimensional cubic lattice and restrict ourselves to half filling and zero temperature. We use a slave-boson method in which the approximation of Gutzwiller, Brinkman, and Rice is recovered on the saddle-point level, but which at the same time is open to various generalizations such as the inclusion of antiferromagnetic long-range order. The slave-boson method yields a ground-state energy in good agreement with quantum Monte Carlo simulations including antiferromagnetic order and spiral states , or with exact diagonalization studies . It has also been used to determine a magnetic phase diagram and to go beyond the Hartree-Fock approximation in problems involving complicated spatial structures such as stripes . In the Kotliar-Ruckenstein approach, two aspects of a physical electron are separated: that it is a fermion and that it affects the occupancy of some site. The first aspect is taken into account by a fermionic field $`f_{i\sigma }`$, while the possible occupancies of the sites are described by bosonic fields: $`e_i`$ describes empty, $`p_{i\sigma }`$ singly occupied, and $`d_i`$ doubly occupied sites. The physical electron field is represented as $`c_{i\sigma }=\stackrel{~}{z}_{i\sigma }f_{i\sigma }`$ with $`\stackrel{~}{z}_{i\sigma }=(1p_{i\sigma }^+p_{i\sigma }d_i^+d_i)^{1/2}(e_i^+p_{i\sigma }+p_{i\sigma }^+d_i^+)(1e_i^+e_ip_{i\sigma }^+p_{i\sigma })^{1/2}`$, while appropriate constraints eliminate unphysical states . Thus, the problem posed by the Hubbard interaction is shifted to that of keeping track of the backflow of bosonic excitations, $`\stackrel{~}{z}_{i\sigma }^+\stackrel{~}{z}_{j\sigma }`$, accompanying the itinerant fermions, $`f_{i\sigma }^+f_{j\sigma }`$. Proceeding along the lines of Ref. , we first set up the functional-integral representation of the Hubbard model in terms of the above-mentioned auxiliary fields, integrate out the fermions, and solve the remaining problem in the saddle-point approximation. To describe antiferromagnetism, we divide the lattice into two sublattices, $`A`$ and $`B`$, and look for solutions satisfying the following relations between the sublattice Bose fields: $`e_B=e_A`$, $`p_{B\sigma }=p_{A\sigma }`$, $`d_B=d_A`$, and $`m=p_A^2p_A^2`$, where $`m`$ is the staggered magnetization. For our result, we need the dispersion relations of the renormalized quasiparticle bands, $$ϵ_{\stackrel{}{k}\eta }[X]=4t^{}\mathrm{cos}k_x\mathrm{cos}k_y+\eta t\sqrt{X^2+4(\mathrm{cos}k_x+\mathrm{cos}k_y)^2},$$ (2) where the lattice spacing has been set equal to one, $`X`$ is some dynamically generated staggered magnetic field, and $`\eta =\pm 1`$. The equations for the density per site (which at half filling is equal to one) and the staggered magnetization, $`1`$ $`=`$ $`{\displaystyle \frac{1}{N}}{\displaystyle \underset{\stackrel{}{k}\eta }{}}f(ϵ_{\stackrel{}{k}\eta }[X]\stackrel{~}{\mu }),`$ (3) $`m(X,\stackrel{~}{\mu })`$ $`=`$ $`{\displaystyle \frac{1}{N}}{\displaystyle \underset{\stackrel{}{k}}{}}{\displaystyle \frac{X}{\sqrt{X^2+4(\mathrm{cos}k_x+\mathrm{cos}k_y)^2}}}`$ (5) $`\times \left[f(ϵ_\stackrel{}{k}[X]\stackrel{~}{\mu })f(ϵ_{\stackrel{}{k}+}[X]\stackrel{~}{\mu })\right],`$ can be solved unambiguously for $`X`$ and the effective chemical potential $`\stackrel{~}{\mu }`$, to yield functions $`\stackrel{~}{\mu }(m)`$ and $`X(m)`$. In Eqs. (3) and (5), $`N`$ is the total number of lattice sites, the sum is over the first Brillouin zone, and $`f(ϵ)=\mathrm{\Theta }(ϵ)`$ is the Fermi function at zero temperature. From the mean-field equation $$\frac{q(m,d^2)}{d^2}K(m)+U=0,$$ (6) where the functions $`q(m,d^2)`$ and $`K(m)`$ are given by $`q(m,d^2)`$ $`=`$ $`{\displaystyle \frac{4d^2}{1m^2}}\left[12d^2+\sqrt{(12d^2)^2m^2}\right],`$ (7) $`K(m)`$ $``$ $`{\displaystyle \frac{1}{N}}{\displaystyle \underset{\stackrel{}{k}\eta }{}}ϵ_{\stackrel{}{k}\eta }[X(m)]f(ϵ_{\stackrel{}{k}\eta }[X(m)]\stackrel{~}{\mu }(m))`$ (9) $`+tmX(m),`$ we obtain the average portion of doubly occupied sites as a function of the staggered magnetization, $`d^2(m)`$. This function along with Eqs. (7) and (9) allow to write the ground-state energy per site as a function of the staggered magnetization as $$e(m)=q(m)K(m)+Ud^2(m).$$ (10) This result has an intuitive interpretation: $`K(m)`$ is the kinetic energy of noninteracting lattice fermions with nearest and next-nearest neighbor hopping, subject to an internal staggered magnetic field $`tX(m)`$. The renormalization factor $`q(m)`$ accounts for the reduction of the hopping amplitudes due to the local correlations and is characteristic of the Gutzwiller approximation. In our scheme, $`q`$ arises from the expectation value $`\stackrel{~}{z}_{i\sigma }^+\stackrel{~}{z}_{j\sigma }`$ and thus represents the average effect of the backflowing slave bosons. The second term in Eq. (10) is the contribution of the Hubbard interaction to the energy. For $`U=0`$, Eqs. (6) and (7) yield $`d^2=(1m^2)/4`$ and $`q=1`$, so $`ϵ(m)=K(m)`$. For $`t^{}=m=0`$, Eqs. (2)-(9) imply: $`X=\stackrel{~}{\mu }=0`$; $`K=2_{\mathrm{}}^0𝑑ϵD_0(ϵ)ϵ`$, where $`D_0(ϵ)`$ is the density of states for noninteracting electrons; $`d^2=\frac{1}{4}(1\frac{U}{U_c})`$ with $`U_c=8|K|`$; and $`q=1(\frac{U}{U_c})^2`$. We thus recover a result of Ref. . In the strong-correlation limit, $`Ut`$, we find up to leading order in $`t/U`$: $`d^2=4t^2/U^2`$, $`q=14t^2/U^2`$, $`K=8t^2/U`$, and $`m=18t^2/U^2`$. Hence, the ground-state energy is $`e=4t^2/U`$. This result is qualitatively correct: In the strong-correlation limit, the half-filled Hubbard model can be mapped onto the Heisenberg antiferromagnet with an exchange coupling constant $`J4t^2/U`$. If we treat the electron spins as Ising spins and use that the coordination number of our two-dimensional lattice is four, we also obtain $`e=J`$. Therfore, the mean-field theory gives rise to the energy scale $`J`$ and correctly accounts for a ground-state energy of the order of $`J`$ per site. Finally, we note that the leading $`1/U`$ corrections do not depend on $`t^{}`$. More generally, the entire insulating phase is unaffected by next-nearest-neighbor hops. $`t^{}`$ enters only via Eqs. (3), (5) and (9). For the insulator, however, only the band $`ϵ_\stackrel{}{k}`$ — but the entire one — contributes to the $`\stackrel{}{k}`$ sums in these equations, and $`t^{}`$ drops out. The ground state for given model parameters $`U`$ and $`\alpha t^{}/t`$ corresponds to the minimum of the energy function (10). $`\alpha `$ is a measure for the degree of magnetic frustration and is varied from zero to one. We may restrict ourselves to positive $`\alpha `$, since a sign change of $`\alpha `$ is tantamount to a particle-hole transformation, $`a_{i\eta }\mathrm{exp}\{i\stackrel{}{Q}\stackrel{}{R}_i\}a_{i|\eta }^+`$ ($`a`$ and $`a^+`$ describe the quasiparticles with dispersion (2), $`\stackrel{}{Q}=(\pi ,\pi )`$, and $`\stackrel{}{R}_i`$ is the lattice vector to site $`i`$). The evolution of $`e(m)`$ as a function of $`U`$ and $`\alpha `$ reveals how the transitions between the various phases take place. For $`\alpha 0`$, The inverse susceptibility, $`\chi ^1`$, is known explicitly and changes its sign from positive to negative as $`U`$ is increased to above $`U_08|K(0)|(\sqrt{|K(0)|/(|K(0)|t\gamma )}1)`$. Here, $`1/\gamma (1/N)_\stackrel{}{k}^{}1/|\mathrm{cos}k_x+\mathrm{cos}k_y|`$ and $`_\stackrel{}{k}^{}`$ denotes the sum over those regions of the first Brillouin zone which are restricted by the condition $`ϵ_\stackrel{}{k}(X=0)<\stackrel{~}{\mu }(m=0)<ϵ_{\stackrel{}{k}+}(X=0)`$. $`U_0`$ and $`\gamma `$ depend only on $`\alpha `$. Whether the system is metallic or insulating depends on the value of the ground-state magnetization: If it exceeds a certain value, $`m_{\text{MIT}}`$, a gap opens up in the single-particle spectrum and the system goes insulating. This can be seen from Eq. (2) if we use that $`m`$ increases monotonically as a function of $`X`$. Consequently, the insulator is always antiferromagnetically ordered. We infer from Eqs. (2)-(5) that $`m_{\text{MIT}}`$ does not depend on $`U`$, but on the level of magnetic frustration: Due to perfect nesting, $`lim_{\alpha 0}m_{\text{MIT}}=0`$. As $`\alpha `$ is turned on, $`m_{\text{MIT}}`$ increases monotonically as a function of $`\alpha `$. In our numerical investigation, we used a tetrahedron method and, for $`\alpha <0.2`$ and $`m<m_{\text{MIT}}`$, up to $`10^4\times 10^4`$ points to do the $`\stackrel{}{k}`$ sums in Eqs. (3), (5) and (9). In the discussion of our numerical results, we must distinguish between three regimes of magnetic frustration and may restrict the discussion of $`e(m)`$ to positive magnetizations. Fig. 1 illustrates how $`e(m)`$ evolves as a function of $`U`$ and $`\alpha `$. The resulting phase diagram is displayed in Fig. 2. For small levels of magnetic frustration, $`0<\alpha <0.06`$ (first column of Fig. 1), we first find a second-order transition from the paramagnetic to the antiferromagnetic metal at a critical value $`U_{\text{mag}}=U_0`$. Upon further increasing $`U`$, the resulting minimum is continuously shifted towards higher magnetizations until it crosses $`m_{\text{MIT}}`$ at a second critical value, $`U_{\text{MIT}}`$. Consequently, the metal-insulator transition is also of second order. Since $`m_{\text{MIT}}`$ vanishes as $`\alpha 0`$, both transitions coincide in this limit, $`U_{\text{mag}}=U_{\text{MIT}}=U_0`$. We also know that $`U_0=0`$ if $`\alpha =0`$. For intermediate levels of magnetic frustration, $`0.06\alpha <0.14`$ (middle column of Fig. 1), the magnetic transition is still of second order and occurs at $`U_0`$, $`U_{\text{mag}}=U_0`$. On the other hand, $`m_{\text{MIT}}`$ is sufficiently large for the metal-insulator transition to take place differently: Before the ground-state magnetization of the antiferromagnetic metal reaches $`m_{\text{MIT}}`$ upon increasing $`U`$, a second minimum at a magnetization above $`m_{\text{MIT}}`$ has emerged and become the absolute minimum of $`e(m)`$. Consequently, the metal-insulator transition is now of first order. The transition lines $`U_{\text{MIT}}(\alpha )`$ from the small-$`\alpha `$ and intermediate-$`\alpha `$ regimes meet at $`\alpha =0.06`$ (filled circle in Fig. 2). At this point, the two degenerate minima of the first-order transition merge at $`m_{\text{MIT}}`$. Finally, the antiferromagnetic metallic phase disappears gradually as $`\alpha 0.14`$ (filled square in Fig. 2). For large degrees of magnetic frustration, $`\alpha 0.14`$ (right column of Fig. 1), the antiferromagnetic and metal-insulator transitions coincide, $`U_{\text{MIT}}=U_{\text{mag}}`$, because the second-order transition at $`U_0`$ is now preempted by the first-order one: By the time the minimum at $`m=0`$ bifurcates, the one above $`m_{\text{MIT}}`$ has already evolved into the absolute one, and remains to be so, as $`U`$ is further increased. We have also considered the Hubbard model in the Hartree-Fock approximation. Earlier work on the $`tt^{}`$ Hubbard model in Hartree-Fock approximation was carried out in Refs. , , and . As for the two-dimensional half-filled case, these works did not conclusively answer whether an antiferromagnetic metal is stable in a certain parameter regime. Within our formalism, the Hartree-Fock approximation turns out to be tantamount to taking $`q=1`$ and $`d^2(m)=(1m^2)/4`$ in Eq. (10), while the function $`K(m)`$ is again determined by Eqs. (3), (5), and (9). As a consequence, $`\chi ^1=2(\gamma tU/4)`$, which changes its sign at $`U_0(\alpha )=4t\gamma (\alpha )`$. This is the small-$`\alpha `$ limit of the corresponding mean-field expression. Expanding $`e^{\text{HF}}(m)`$ one step further yields the same $`|m|^3\mathrm{log}|m|`$ term as in mean-field theory, but with $`q=1`$. This suggests that the magnetic transition is of second order and takes place at $`U_0(\alpha )`$. In fact, our numerical investigation of the evolution of $`e^{\text{HF}}(m)`$ as $`U`$ is increased reveals the following behavior for all finite values of $`\alpha `$: After a second-order transition from the paramagnetic to the antiferromagnetic metal at $`U_{\text{mag}}=U_0(\alpha )`$, the minimum of $`e^{\text{HF}}(m)`$ is rapidly but continuously displaced towards higher magnetizations, until it exceeds $`m_{\text{MIT}}`$. Thus, the metal-insulator-transition is also of second order. Fig. 3 displays the phase diagram in Hartree-Fock approximation. The error bars of the transition lines $`U_{\text{mag}}(\alpha )`$ and $`U_{\text{MIT}}(\alpha )`$ become of the same order than $`U_{\text{MIT}}(\alpha )U_{\text{mag}}(\alpha )`$ for $`\alpha 0.001`$, and our numerics is trustworthy down to about $`\alpha 0.002`$. In summary, we have investigated the effect of magnetic frustration on the metal-insulator transition in the two-dimensional half-filled Hubbard model within a slave-boson approach, and we have compared our results to the Hartree-Fock approximation. Within the slave-boson mean-field theory, our main results are: First, magnetic frustration helps stabilize an antiferromagnetic metal for $`t^{}/t0.14`$. Second, for $`t^{}/t0.06`$, the metal-insulator transition is of first order. Finally, all other transitions between the various phases are of second order. By contrast, in Hartree-Fock approximation, the magnetic and metal-insulator transitions are always separate and of second order. Both in Hartree-Fock approximation and in the slave-boson mean-field theory, the insulator is always antiferromagnetically ordered. We gratefully acknowledge discussions with R. Frésard and R. Chitra. This work was supported by the NSF DMR 95-29138. E.L. is partly funded by the Deutsche Forschungsgemeinschaft.
no-problem/9908/astro-ph9908069.html
ar5iv
text
# 1 ISO light curves. Dotted lines are the mean flux values of the best sampled period. ISO observations of the BL Lac object PKS 2155–304 E. Bertone<sup>1</sup>, G. Tagliaferri<sup>1</sup>, G. Ghisellini<sup>1</sup>, A. Treves<sup>2</sup>, L. Chiappetti<sup>3</sup>, L. Maraschi<sup>1</sup> <sup>1</sup> Osservatorio Astronomico di Brera, via Brera 28, I–20121 Milano, Italy – email: bertone@merate.mi.astro.it <sup>2</sup> Università dell’Insubria, via Lucini 3, I–22100 Como, Italy <sup>3</sup> IFCTR/CNR, via Bassini 15, I–20133 Milano, Italy 1. Introduction The Infrared Space Observatory<sup>1</sup><sup>1</sup>1ISO is an ESA project with instruments funded by ESA Member States (especially the PI countries: France, Germany, the Netherlands and the United Kingdom) with the participation of ISAS and NASA. (ISO) observed the BL Lac object PKS 2155–304 fifteen times, from May 7 to June 8 1996. Twelve observations were carried out in a best sampled period of 15 days. Two additional observations were performed on 1996 November 23 and 1997 May 15. These observations were part of a multiwavelength monitoring which involved UV, X–ray, $`\gamma `$–ray satellites and ground based telescopes. The instruments used were the ISOCAM camera in the range 2.5 – 18 $`\mu `$m and the ISOPHOT photometer, from 25 up to 170 $`\mu `$m. 2. The ISO light curves The two ISOCAM light curves, at 4.0 and 14.3, $`\mu `$m and the two ISOPHOT light curves, at 60 and 90 $`\mu `$m, are shown in Fig. 1. There is no evidence of time variability of the flux at the four wavelengths, but the large errors (more than 10% for the camera fluxes and up to 50% for the photometer data) can hide smaller variations. We calculated the mean relative error and we obtained 3 sigma limits for the lowest detectable flux variations of 32%, 36%, 76% and 132% at 4.0, 14.3, 60 and 90 $`\mu `$m respectively. The mean flux values of the best sampled period, from 1996 May 13 to May 27, are $`54.8\pm 1.8`$ mJy at 4.0 $`\mu `$m, $`90.8\pm 3.2`$ mJy at 14.3 $`\mu `$m, $`315\pm 27`$ mJy at 60 $`\mu `$m and $`250\pm 34`$ mJy at 90 $`\mu `$m. 3. The ISO spectrum The infrared spectrum of PKS 2155–304 was sampled, using 16 broad band filters, from 2.8 to 170 $`\mu `$m (Fig. 2a). It is well fitted with a single power law of energy spectral index $`\alpha =0.40\pm 0.06`$. This power–law shape strongly supports the hypothesis that the far– and mid– infrared emission of this source is entirely generated by synchrotron process, excluding, therefore, important contributions from thermal sources. The emission of the host galaxy of PKS 2155–304, a big elliptical that is seen in NIR images (Kotilainen et al. 1998), is negligible at longer IR wavelengths (Bertone et al. 1999). From the SED of the source in Fig. 2b, which shows the simultaneous data of 1996 May, one can notice that the ISO data lie well on the extrapolation between the radio and the optical data, and follow well the shape of the synchrotron peak, supporting the fact that the synchrotron emission is dominant in the ISO band. References Bertone E., Tagliaferri G., Ghisellini G., et al., 1999, in preparation Kotilainen J.K., Falomo R. & Scarpa R., 1998, A&A 336, 479 Urry C.M., Sambruna R.M., Brinkmann W.P. & Marshall H., 1998, in: Scarsi L., Bradt H., Giommi P., Fiore F. (eds.) The Active X-Ray Sky. Nucl. Phys. B (Proc. Suppl.) 69, 419
no-problem/9908/cond-mat9908094.html
ar5iv
text
# The influence of correlated hopping on valence and metal-insulator transitions in the Falicov-Kimball model ## Abstract The extrapolation of small-cluster exact-diagonalization calculations is used to examine the influence of correlated hopping on valence and metal-insulator transitions in the one-dimensional Falicov-Kimball model. It is shown that the ground-state phase diagram as well as the picture of valence and metal-insulator transitions found for the conventional Falicov-Kimball model (without correlated hopping) are strongly changed when the correlated hopping term is added. The effect of correlated hopping is so strong that it can induce the insulator-metal transition, even in the half-filled band case. PACS nrs.:75.10.Lp, 71.27.+a, 71.28.+d, 71.30.+h In the past decade, a considerable amount of effort has been devoted to understanding of the multitude of anomalous physical properties of rare-earth and transition-metal compounds. Recent theoretical works based on exact numerical and analytical calculations showed that many of these anomalous features, e.g., mixed valence phenomena, metal-insulator transitions, etc., can be describe very well within different versions of the Falicov-Kimball model (FKM) . For example, it was found that the spinless FKM, in the pressure induced case, can describe both types of intermediate-valence transitions observed experimentally in rare-earth compounds: a discontinuous insulator-insulator transition for sufficiently strong interactions and a discontinuous insulator-metal transition for weak interactions . In addition, at nonzero temperatures this model is able to provide the qualitative explanation for anomalous large values of the specific heat coefficient and for extremely large changes of electrical conductivity found in some intermediate-valence compounds (e.g., in SmB<sub>6</sub>). Moreover, very recently the spin-one-half version of the FKM has been used successfully for a description of a discontinuous intermediate-valence transition (accompanied by a discontinuous insulator-metal transition) in SmS as well as for a description of an anomalous magnetic response of the Yb-based valence-fluctuating compounds . These results show that the FKM could, in principle, yield the correct physics for describing rare-earth compounds. On the other hand it should be noted that the conventional FKM neglects all nonlocal interaction terms and thus it is questionable whether above mentioned results persist also in more realistic situations when nonlocal interactions will be turned on. An important nonlocal interaction term obvious absent in the conventional FKM is the term of correlated hopping in which the $`d`$-electron hopping amplitudes $`t_{ij}`$ between neighbouring lattice sites $`i`$ and $`j`$ depend explicitly on the occupancy $`(f_i^+f_i)`$ of the $`f`$-electron orbitals. To examine effects of this term on valence and metal-insulator transitions in the FKM we adopt here the following general form for the nearest-neighbour matrix elements $$\stackrel{~}{t}_{ij}=t_{ij}+t_{ij}^{}(f_i^+f_i+f_j^+f_j),$$ (1) which represent a much more realistic type of electron hopping than the conventional hopping ($`t_{ij}=t`$ if $`i`$ and $`j`$ are the nearest neighbours, and zero otherwise). Thus the spinless FKM in which the effects of correlated hopping are included can be written as $$H=\underset{<ij>}{}t_{ij}d_i^+d_j+\underset{<ij>}{}t_{ij}^{}(f_i^+f_i+f_j^+f_j)d_i^+d_j+U\underset{i}{}f_i^+f_id_i^+d_i+E_f\underset{i}{}f_i^+f_i,$$ (2) where $`f_i^+`$, $`f_i`$ are the creation and annihilation operators for an electron in the localized state at lattice site $`i`$ with binding energy $`E_f`$ and $`d_i^+`$, $`d_i`$ are the creation and annihilation operators of the itinerant spinless electrons in the $`d`$-band Wannier state at site $`i`$. The first term of (2) is the kinetic energy corresponding to quantum mechanical hopping of the itinerant $`d`$-electrons between the nearest-neighbour sites $`i`$ and $`j`$. The second term is just the correlated hopping term discussed above. The third term is the on-site Coulomb interaction between the $`d`$-band electrons with density $`n_d=\frac{1}{L}_id_i^+d_i`$ and the localized $`f`$-electrons with density $`n_f=\frac{1}{L}_iw_i`$, where $`L`$ is the number of lattice sites. The last term stands for the localized $`f`$ electrons whose sharp energy level is $`E_f`$. Since in this spinless version of the FKM without hybridization the $`f`$-electron occupation number $`f_i^+f_i`$ of each site $`i`$ commutes with the Hamiltonian (2), the $`f`$-electron occupation number is a good quantum number, taking only two values: $`w_i=1`$ or 0, according to whether or not the site $`i`$ is occupied by the localized $`f`$ electron. Therefore the Hamiltonian (2) can be written as $$H=\underset{<ij>}{}h_{ij}(w)d_i^+d_j+E_f\underset{i}{}w_i,$$ (3) where $`h_{ij}(w)=\stackrel{~}{t}_{ij}(w)+Uw_i\delta _{ij}`$ and $$\stackrel{~}{t}_{ij}(w)=t_{ij}+t_{ij}^{}(w_i+w_j).$$ (4) Thus for a given $`f`$-electron configuration $`w=\{w_1,w_2\mathrm{}w_L\}`$ defined on a one-dimensional lattice with periodic boundary conditions, the Hamiltonian (2) is the second-quantized version of the single-particle Hamiltonian $`h(w)`$, so the investigation of the model (3) is reduced to the investigation of the spectrum of $`h`$ for different configurations of $`f`$ electrons. Since the $`d`$ electrons do not interact among themselves, the numerical calculations precede directly in the following steps (in the next we consider only the case $`N_f+N_d=L`$, which is the point of the special interest for the mixed-valence phenomena). (i) Having $`w=\{w_1,w_2\mathrm{}w_L\}`$, $`U`$, $`E_f`$ and the nearest-neighbour hopping amplitudes $`t`$ and $`t^{}`$ fixed, (in the following $`t=1`$ and all energies are measured in units of $`t`$) find all eigenvalues $`\lambda _k`$ of $`h(w)`$. (ii) For a given $`N_f=_iw_i`$ determine the ground-state energy $`E(w,U,E_f)=_{k=1}^{LN_f}\lambda _k+E_fN_f`$ of a particular $`f`$-electron configuration $`w`$ by filling in the lowest $`N_d=LN_f`$ one-electron levels. (iii) Find the $`w^0`$ for which $`E(w,U,E_f)`$ has a minimum. Repeating this procedure for different values of $`U,t^{}`$ and $`E_f`$, one can study directly the the ground-state phase diagram and valence transitions (a dependence of the $`f`$-electron occupation number on the $`f`$-level position $`E_f`$) in the FKM with correlated hopping. The most interesting question that arises for the FKM with correlated hopping is whether the correlated hopping term can change the ground state phase diagram and the picture of valence and metal-insulator transitions found for the conventional FKM ($`t=1`$ and $`t^{}=0`$). The nature of the ground state, its energetic and structural properties and the correlation-induced metal-insulator transitions are subjects of special interest. For the conventional FKM these problems are well understood at least at half-filling ($`E_f=0,n_f=n_d=1/2`$). In this case the localized $`f`$-electrons fill up one of two sublattices of the hypercubic lattice (the charge density wave state) and the corresponding ground state is insulating for all $`U>0`$. Thus, for the finite interaction strength there is no correlation-induced metal-insulator transition in the half-filled band case. Outside half-filling the situation is more complicated since there exist the analytical results only for a restricted class of configurations and exact-numerical results are limited to small finite clusters . In spite of these restrictions both analytical and numerical results yield the same ground-state phase diagram which consists of only two main domains: the domain of the phase separation and the domain of the most homogeneous configurations. The domain of the phase separation is restricted to a small region of low $`f`$-electron concentrations ($`n_f<1/4`$) and weak Coulomb interactions ($`U<1.2`$). In this domain the ground state are incoherent (metallic) mixtures of the empty configuration (whose length $`l`$ is at least $`L/2`$) and a configuration $`w`$ (whose length is $`l<L/2`$). Outside this domain the ground states are the most homogeneous configurations, which are insulating for all $`f`$-electron concentrations. The boundary between these two domains is the boundary of correlation-induced metal-insulator transitions. One can expect, on the base of simple arguments, that the ground-state phase diagram of the FKM with correlated hopping will be fully different from one discussed above for the conventional FKM. Indeed, the following selection of hopping matrix amplitudes $`t=1`$ and $`t^{}>0`$ may favor the segregated configuration $`\{11\mathrm{}100\mathrm{}0\}`$ since the itinerant $`d`$ electrons have the lower kinetic energy in this state. This mechanism could lead, for example, to the instability of the CDW state $`\{10\mathrm{}10\}`$ that is the ground state for $`t^{}=0`$ and thereby to a metal-insulator transition, even in the half-filled band case. To verify this conjecture we have performed an exhaustive study of the model on finite clusters (up to 32 sites) for a wide range of parameters $`t^{}`$ and $`U`$. The results of numerical calculations are summarized in Fig. 1 in the form of the $`t^{}`$-$`U`$ phase diagram. In addition to the alternating configuration $`w_1=\{10\mathrm{}10\}`$ that is the ground state at $`t^{}=0`$ for all nonzero $`U`$ we found two new phases that can be the ground states of the model, and namely, the alternating configuration $`w_2=\{1100\mathrm{}1100\}`$ with double period and the segregated configuration $`w_3=\{11\mathrm{}100\mathrm{}0\}`$. Thus at nonzero $`t^{}`$ the configuration $`w_1`$ becomes unstable against the transition to $`w_2`$ and $`w_3`$. The transition from $`w_1`$ to $`w_2`$ is the insulator-insulator transition since both $`w_1`$ and $`w_2`$ configurations have the finite gaps $`\mathrm{\Delta }`$ at the Fermi energy for nonzero values of $`U`$ (see the inset in Fig. 1). On the other hand the segregated configuration $`w_3`$ is metallic ($`\mathrm{\Delta }=0`$) for all nonzero $`U`$ and so the transition from $`w_1`$ to $`w_3`$ (as well as from $`w_2`$ to $`w_3`$) is the insulator-metal transition. Thus we arrive to the very important conclusion, and namely, that the correlated hopping term can induce the insulator-metal transition, even in the half-filled band case. The fact that the correlated hopping plays the important role in stabilizing the ground state indicates that the picture of valence and metal-insulator transitions found in our previous papers within the conventional FKM could be dramatically changed if finite values of $`t^{}`$ will be considered. To verify this conjecture we have performed the study of the model for two representative values of $`U`$ ($`U=1.5`$ and $`U=4`$) and a selected set of $`t^{}`$ values. The results of numerical calculations obtained for $`U=1.5`$ on finite clusters (up to 32 sites) can be summarized as follows: (i) For sufficiently small values of $`t^{}`$ ($`t^{}0.3`$) the ground states are the most homogeneous configurations for all $`f`$-electron concentrations. (ii) For intermediate values of $`t^{}`$ ($`t^{}0.6`$) the ground states are the segregated configurations for $`n_f<0.5`$ and the most homogeneous configurations for $`n_f0.5`$. (iii) For large values of $`t^{}`$ ($`t^{}>0.8`$) the ground states are only the segregated configurations. Since no significant finite-size effects have been observed on clusters up to 32 sites we suppose that these results can be satisfactory extrapolated on large systems and used for a study of effects of correlated hopping on valence and metal-insulator transitions. Typical valence transitions obtained on the extrapolated set of configurations are presented in Fig. 2. For $`t^{}=0.3`$ the valence transition holds all characteristic features found in our previous papers for $`t^{}=0`$. The basic structure of the transition is formed by the most homogeneous configurations with the smallest periods ($`n_f=1/5,1/4,1/3,1/2,2/3,3/4,4/5`$). This primary structure is independent of $`L`$ while the secondary structure corresponding to remaining configurations strongly depends on $`L`$ and forms the gradual transition between two valence states from the primary structure. The correlated hopping term reduces considerably the width of the primary structure, especially for $`n_f0.5`$, but it does not change the nature of the ground state. The ground states for $`t^{}=0.3`$ are the most homogeneous configurations which are insulating (see the inset in Fig. 2) and thus there is no insulator-metal transition induced by external pressure for small values of $`t^{}`$. (It is generally supposed that in the pressure induced case $`E_f`$ is proportional to pressure.) From this point of view the situation looks more hopeful for intermediate values of $`t^{}`$. As was mentioned above for this case the ground states are the segregated (metallic) configurations (for $`n_f<0.5`$) and thus one could expect the pressure induced insulator-metal transition at $`E_f=E_c`$ (corresponding to $`n_f=n_c=0.5`$), of course if the segregated configurations persist as the ground states on some finite region of $`E_f`$ values. Fig. 2 illustrates that such a situation indeed becomes. For $`t^{}=0.6`$ the segregated configurations are the ground states in the wide region above $`E_c=0.02`$ and the valence transition in this region is continuous. Below $`E_c`$ the ground states are the most homogeneous (insulating) configurations. In this region the primary structure is practically fully suppressed and only the intermediate valence states with $`n_f=1/2`$ and $`n_f=2/3`$ have the finite regions of stability. At $`E_f=E_c`$ the FKM exhibits a discontinuous insulator-metal transition that is apparently induced by correlated hopping. This result shows how crucial role plays the correlated hopping in the mechanism of valence and insulator-metal transitions. Already relatively small values of $`t^{}`$ can fully destroy the picture of valence and metal-insulator transitions found for the conventional FKM and therefore the correlated hopping term cannot be neglected in the correct description of valence and metal-insulator transitions. Of course, for large values of $`t^{}`$ we can expect the largest deviations from $`t^{}=0`$ case, since the ground states in this region are only the segregated configurations which contrary to the most homogeneous configurations are metallic. The inset in Fig. 2 confirms this conjecture. The valence transitions for large values of $`t^{}`$ are only continuous and fully differ from $`t^{}=0`$ case. In the strong coupling region ($`U=4`$) the ground states are either the most homogeneous configurations (for small and intermediate $`t^{}`$) or the segregated configurations (for large $`t^{}`$). Thus for small and intermediate values of $`t^{}`$ the strong coupling picture of valence and metal-insulator transitions will be the same as one for $`t^{}=0`$. All ground states are insulating and the valence transitions, for both conventional FKM and FKM with correlated hopping consist of a few discontinuous valence transitions which number is reduced with increasing $`U`$. Only one difference between valence transitions obtained within these two models is that the width of the valence transition in the FKM with correlated hopping is reduced due to finite $`t^{}`$. Fig. 3 illustrates that the effect is very strong even for relatively small values of $`t^{}`$. This indicates that the correlated hopping plays important role also in the strong coupling limit although unlike the weak-coupling limit now small values of $`t^{}`$ do not change qualitatively the picture of valence and metal-insulator transitions. In order to change qualitatively the picture of valence and metal-insulator transitions in the strong coupling limit, larger values of $`t^{}`$ should be considered. For example, we have found that for $`U=4`$ and $`t^{}=1.2`$ the ground states are the segregated (metallic) configurations for all $`f`$-electron concentrations and the corresponding valence transition is continuous. The exactly opposite effect on valence transitions in the strong coupling limit has the correlated hopping term with negative values of $`t^{}`$. As shown in Fig. 4. the correlated hopping with $`t^{}<0`$ stabilizes the insulating intermediate-valence states corresponding to the most homogeneous configurations with the smallest periods (which are the ground states in this region) and this effect is very strong even for relatively small values of $`t^{}`$. This again leads to conclusion that the correlated hopping must be taken into account in the correct description of valence and metal-insulator transitions. This work was supported by the Slovak Grant Agency VEGA under grant No. 2/4177/97. Numerical results were obtained using computational resources of the Computing Centre of the Slovak Academy of Sciences. Figure Captions Fig. $`1.`$ $`t^{}`$-$`U`$ phase diagram of the FKM with correlated hopping at half-filling ($`E_f=0,n_f=n_d=0.5`$). Three different phases correspond to the alternating configuration $`w_1=\{10\mathrm{}10\}`$, the alternating configuration $`w_2=\{1100\mathrm{}1100\}`$ with double period and the segregated configuration $`w_3=\{1\mathrm{}10\mathrm{}0\}`$. For $`L=24`$ the phase diagram has been obtained over the full set of $`f`$-electron configurations, while for $`L=600`$ only the restricted set of configurations $`w_1,w_2`$ and $`w_3`$ has been used to determine the phase boundaries. The inset shows $`t^{}`$-dependence of energy gaps corresponding to the ground state configurations for different values of $`U`$. Fig. $`2.`$ Dependence of the $`f`$-electron occupation number $`n_f`$ on the $`f`$-level position $`E_f`$ for different values of correlated hopping $`t^{}`$ at $`U=1.5`$. The upper inset shows energy gaps corresponding to the most homogeneous configurations that are ground states for $`t^{}=0.3`$ (for all $`f`$-electron concentrations) and $`t^{}=0.6`$ (for $`n_f0.5`$). The lower inset shows examples of $`n_f(E_f)`$ behaviour for large values of $`t^{}`$. Fig. $`3.`$ Dependence of the $`f`$-electron occupation number $`n_f`$ on the $`f`$-level position $`E_f`$ for different values of correlated hopping $`t^{}`$ at $`U=4`$. The inset shows energy gaps corresponding to the most homogeneous configurations that are ground states for all $`f`$-electron concentrations for $`t^{}=0.4`$ and $`U=4`$ Fig. $`4.`$ Dependence of the $`f`$-electron occupation number $`n_f`$ on the $`f`$-level position $`E_f`$ for different values of correlated hopping $`t^{}0`$ at $`U=4`$. The inset shows energy gaps corresponding to the most homogeneous configurations that are ground states for all $`f`$-electron concentrations for $`t^{}=0.2`$ and $`t^{}=0.4`$.
no-problem/9908/cond-mat9908025.html
ar5iv
text
# Critical-point finite-size scaling in the microcanonical ensemble ## I Introduction Statistical mechanics can be formulated in any of a set of ensembles distinguished by the relationship between the system and its environment . The principal members of this set are the microcanonical (prescribed energy) and canonical (prescribed temperature) ensembles. In the thermodynamic limit (when it exists) the ensembles yield the same predictions (and are, in this sense, equivalent) and the choice of ensemble is a matter of practical convenience. The canonical ensemble tends to win this contest because it circumnavigates the hard-constant-energy constraint imposed by the microcanonical ensemble. The two ensembles are, however, not always equivalent . They differ for systems which are ‘small’ in some sense: inherently small systems such as nuclei or clusters ; systems with unscreened long-range forces where the thermodynamic limit is problematic; and systems at critical points , which are our principal concern here. Theoretical studies of critical phenomena are almost invariably conducted within the framework of the canonical ensemble . In consequence there is no substantive framework within which to interpret computational studies of microcanonical critical behavior. Such studies do, nevertheless, exist, having been motivated, variously, by the belief that the microcanonical framework may have some computational advantages and by the discovery that, apparently, critical anomalies in the microcanonical heat capacity are significantly enhanced with respect to their canonical counterparts. This paper goes some way towards supplying the missing framework. We develop (section 2) a finite-size-scaling theory for the microcanonical entropy (the density of states) of a system with a critically-divergent heat capacity. In so doing we have, of necessity, to consider more general questions about the structure of the density of states of a finite-size system –in particular the implications of well-established results for the finite-size structure of the canonical partition function . Though somewhat more than a phenomenology, our theory falls short of being microscopically explicit: to determine an explicit form for the relevant scaling function we need to appeal (section 3) to Monte Carlo (MC) measurements of the critical canonical energy probability distribution (pdf). The canonical energy pdf itself has a near-critical finite-size-scaling form which has featured in a number of studies of critical points in fluids and lattice gauge theories . Since energy fluctuations (like the critical anomaly in the canonical specific heat which they control) are relatively weak (by comparison with the fluctuations of the order parameter, and the divergence of its response function) the degree of ‘scaling’ reported in previously measured energy pdfs has been relatively poor –unsatisfactorily so for our purposes here. This problem is addressed in section 3. We show that one can fold out (from the measured distributions) the sub-dominant (but significant) non-scaling effects that are associated with the constant background contribution to the canonical heat capacity, negative in the case of the 3d Ising model . This procedure exposes the underlying behavior, which manifests scaling to an impressive degree. In addition to providing us with the platform needed for this work, this procedure may offer the basis for improving the mixed-scaling-field theory of critical points in systems that belong to the Ising universality class but which do not have full Ising symmetry; recent studies have suggested that the current framework is not fully satisfactory. The scaling form for the critical energy pdf allows us to determine the scaling form of the microcanonical entropy. In section 4 we explore this form and show that it is consistent with predictions for both the bulk-critical limit (as regards the parameters characterizing the specific heat singularity ) and the finite-size critical limit (the Fisher-Privman constant ). The microcanonical entropy also provides us with a unified basis for dealing with both the canonical and the microcanonical specific heats (section 5). We show that the ‘corrections’ to the scaling behavior of the canonical specific heat (the negative background constant) have subtle consequences for the microcanonical behavior. In particular they serve to amplify the difference between microcanonical and canonical behavior, and are at least partially responsible for the strength of the anomaly observed in some microcanonical studies . ### A The microcanonical scaling ansatz We consider a d-dimensional many-body system of linear dimension $`L`$; we assume hypercubic geometry with periodic boundary conditions. The canonical partition function is, in principle, a discrete sum over system microstates ($`r`$) or system energy levels ($`s`$): $$Z(\beta ,L)=\underset{r}{}e^{\beta E_r}=\underset{s}{}\mathrm{\Omega }_se^{\beta E_s}$$ (1a) where $`\mathrm{\Omega }_s`$ is the degeneracy of level $`s`$. We shall suppose that the system is sufficiently large that the sum over levels can be replaced by an integral: $$Z(\beta ,L)=𝑑ϵ\mathrm{\Omega }(ϵ,L)e^{\beta L^dϵ}$$ (1b) where $`ϵE/L^d`$ is the energy density. The function $`\mathrm{\Omega }(ϵ,L)`$ is the density of states; as we have defined it, it is a true density, having dimensions of inverse energy. We note that the transition from the discrete representation (Eq. 1a) to its continuum counterpart (Eq. 1b) requires some care: it is discussed in Appendix A. Our microcanonical scaling theory comprises a proposal for the form of the density of states function. We formulate it in two stages. Consider first a regime remote from critical points or lines of phase coexistence. In such a regime we make the general finite-size ansatz : $$\mathrm{\Omega }(ϵ,L)\left[\frac{L^ds^{\prime \prime }(ϵ)}{2\pi }\right]^{1/2}e^{L^ds(ϵ)}$$ (2) The structure proposed for the prefactor makes this a little more than simply a definition of the microcanonical entropy density $`s(ϵ)`$. In its support we note, first, that one may readily verify it explicitly (Appendix B) in the case of some simple model systems. Secondly we note the implications for the associated canonical partition function. Inserting Eq. (2) into Eq. (1b), a saddle-point integration gives $`Z(\beta ,L)`$ $`=`$ $`\left[{\displaystyle \frac{L^d}{2\pi }}\right]^{1/2}{\displaystyle 𝑑ϵ\left[s^{\prime \prime }(ϵ)\right]^{1/2}e^{L^d\left[s(ϵ)\beta ϵ\right]}}`$ (3) $`=`$ $`e^{L^df(\beta )}\left[1+O(L^d)\right]`$ (4) where $$f(\beta )\beta \widehat{ϵ}s(\widehat{ϵ})$$ (5) and $`\widehat{ϵ}`$ is the solution of $$\beta =s^{}(\widehat{ϵ})$$ (6) Eq. (4) recovers the prefactor-free form of the canonical partition function believed to be widely appropriate in regions (those where the saddle point integration is to be trusted) remote from critical points or lines of phase coexistence . We note that this form is achieved by virtue of the prefactor that does appear in the density of states ansatz (Eq. 2), which is just such as to cancel the contributions made by the fluctuations about the saddle point . The argument we have given leaves open the possibility of power-law corrections to Eq. (4). It has long been believed, and more recently established rather generally , that the corrections to the leading form are actually exponentially small in the system size. Since the saddle-point integration necessarily generates power-law corrections, one must suppose that there are compensating power-law corrections to the ansatz (Eq. 2) for the density of states. This conclusion serves as a warning (already suggested by the double appearance of the function $`s(ϵ)`$ in Eq. (2)) that the microcanonical framework faces problems which are skirted in the canonical formalism . Now, more specifically, consider a system, of the kind specified above, in the vicinity of a critical point. We will suppose that the critical point has a divergent heat capacity; where we need to be more specific we shall assume it is a member of the d=3 Ising universality class (or, more specially still, the d=3 Ising model itself). Within the microcanonical framework the critical point of such a system is located by a critical value $`ϵ_c`$ of the energy density, sharply-defined in the thermodynamic limit. We are concerned with the behavior of the microcanonical entropy for energies in the vicinity of this critical value. To describe this regime we introduce the dimensionless scaling variable $$xa_ϵL^{1/\nu _ϵ}(ϵϵ_c)$$ (7) where $`a_ϵ`$ is an appropriate scale factor and the index is defined by $$\frac{1}{\nu _e}=\frac{1\alpha }{\nu }$$ (8) with $`\alpha `$ the index (assumed positive) characterizing the heat capacity divergence, and $`\nu `$ the correlation length index . We now reformulate and extend our basic ansatz (Eq. 2) with the proposal that, in a region of sufficiently large $`L`$ and sufficiently small $`ϵϵ_c`$ $$\mathrm{\Omega }(ϵ,L)\left[\frac{L^ds^{\prime \prime }(ϵ,L)}{2\pi }\right]^{1/2}e^{L^ds(ϵ,L)}$$ (9a) with $$L^ds(ϵ,L)L^d\left[s_c+\beta _c(ϵϵ_c)\right]+\stackrel{~}{𝒮}(x)$$ (9b) Here $`s_c`$ is an unimportant constant, $`\beta _c`$ is the critical inverse temperature and $`\stackrel{~}{𝒮}(x)`$ is a finite-size-scaling function, universal given some convention on the scale factor $`a_ϵ`$, introduced in Eq. (7). The remainder of this paper is devoted to providing support for this proposal, and exploring the structure of the microcanonical entropy scaling function which it introduces. ## II Determining the scaling function It should be possible to determine the finite-size-scaling function $`\stackrel{~}{𝒮}(x)`$ within the renormalization group framework . We have not done that. Instead we have chosen to learn what we can about this function from its signatures in MC studies of the canonical ensemble. Consider, then, the implications of the scaling form Eq. (9b) for the canonical partition function, Eq. (1b). We suppose initially (we shall have to refine the supposition, shortly) that the relevant part of the energy spectrum is adequately captured by Eq. (9b). Then $$Z(\beta ,L)e^{L^df_0(\beta )}\stackrel{~}{𝒵}(y)$$ (10) where $$f_0(\beta )=\beta ϵ_cs_c$$ (11) and $$\stackrel{~}{𝒵}(y)=𝑑x\sqrt{\frac{\stackrel{~}{𝒮}^{\prime \prime }(x)}{2\pi }}e^{xy+\stackrel{~}{𝒮}(x)}$$ (12) while $$y=a_ϵ^1L^{1/\nu }(\beta \beta _c)$$ (13) provides a scaling measure of the deviation from the critical temperature. We have made use of the hyperscaling relation which links the correlation length index $`\nu `$ and the heat capacity index $`\alpha `$ through $$\frac{1}{\nu }+\frac{1}{\nu _e}=\frac{2\alpha }{\nu }=d$$ (14) The scaling form of the free energy follows: $$F(\beta ,L)\mathrm{ln}Z(\beta ,L)L^df_0(\beta )\mathrm{ln}\stackrel{~}{𝒵}(y)\stackrel{~}{F}(\beta ,L)$$ (15) The canonical energy pdf $$P(ϵ|\beta ,L)Z^1(\beta ,L)\mathrm{\Omega }(ϵ,L)e^{\beta L^dϵ}$$ (16) may also be written in scaling form: $$P(ϵ|\beta ,L)dϵP(x|y,L)dx$$ (17) with $$P(x|y,L)\stackrel{~}{𝒵}^1(y)\sqrt{\frac{\stackrel{~}{𝒮}^{\prime \prime }(x)}{2\pi }}e^{xy+\stackrel{~}{𝒮}(x)}\stackrel{~}{P}(x|y)$$ (18) The scaling predictions for the pdf may be tested by examining its cumulants , for which the free energy is a generator: $$ϵ^{(n)}(\beta ,L)(1)^{n+1}L^{nd}\frac{^nF(\beta ,L)}{\beta ^n}$$ (19) Eq. (15) then implies that the cumulants have the scaling form $$ϵ^{(n)}(\beta ,L)\left[a_ϵL^{1/\nu _ϵ}\right]^n\stackrel{~}{x}^{(n)}(y)+ϵ_c\delta _{n,1}$$ (20) where the scaled cumulants $`\stackrel{~}{x}^{(n)}(y)`$ are universal functions: $$\stackrel{~}{x}^{(n)}(y)=(1)^n\frac{^n\mathrm{ln}\stackrel{~}{𝒵}(y)}{y^n}$$ (21) The canonical mean of the energy density at criticality ($`\beta =\beta _c`$) follows as: $$ϵ_cϵ^{(1)}(\beta _c,L)ϵ_c+\left[a_ϵL^{(1\alpha )/\nu }\right]^1\stackrel{~}{x}^{(1)}(y=0)$$ (22) MC measurements on the 3d Ising model using a range of system sizes (Fig. 1) are fully consistent with this behavior. Eq. (15) implies, likewise, that the canonical variance of the energy density should have the power law behavior $$ϵ^2_cϵ_c^2ϵ^{(2)}(\beta _c,L)a_ϵ^2L^{d+\alpha /\nu }\stackrel{~}{x}^{(2)}(y=0)$$ (23) MC measurements (Fig. 2) are only partially consistent with this prediction: the power law is confirmed, but with an extrapolation whose intercept is far from zero. This inconsistency is reflected in the rather limited success (Fig. 3) of attempts to collapse the measured energy pdfs for different system sizes on to a single scaling form. The source of these problems can be guessed from the implications of Eq. (23) for the canonical specific heat, which it mirrors: the scaling form fails to capture the effects associated with the constant background which constitutes the dominant correction to pure scaling (power-law divergence) of the canonical specific heat. There are two ways to rectify this failure. One might extend the theory to predict the behavior of the (very) finite systems accessible to MC study; or one might seek to correct the MC results to expose the true scaling behavior. We adopt the latter strategy. Define $$\mathrm{\Delta }f(\beta ,L)L^d\left[F(\beta ,L)\stackrel{~}{F}(\beta ,L)\right]$$ (24) the difference between the true free energy density and its asymptotic scaling form (see Eq. (15)). We shall ignore the effects of confluent singularities: they are not the dominant ‘corrections to scaling’ here. Then $`\mathrm{\Delta }f(\beta ,L)`$ is analytic and may be approximated, near $`\beta _c`$, by the expansion $$\mathrm{\Delta }f(\beta ,L)\underset{n=0}{\overset{\mathrm{}}{}}\mathrm{\Delta }f_c^{(n)}\frac{(\beta \beta _c)^n}{n!}$$ (25) These additional contributions to the free energy imply additional contributions to the energy cumulants (Eq. 19): $$\mathrm{\Delta }ϵ^{(n)}(\beta ,L)(1)^{n+1}L^{(n1)d}\frac{^n\mathrm{\Delta }f(\beta ,L)}{\beta ^n}$$ (26) At criticality Eq. (20) must then be modified to read $`ϵ^{(n)}(\beta _c,L)`$ $`=`$ $`\left[a_ϵL^{1/\nu _ϵ}\right]^n\left[\stackrel{~}{x}^{(n)}(y=0)+\mathrm{\Delta }x_c^{(n)}(L)\right]+ϵ_c\delta _{n,1}`$ (27) $``$ $`\left[a_ϵL^{1/\nu _ϵ}\right]^nx^{(n)}(y=0,L)+ϵ_c\delta _{n,1}`$ (28) where $$\mathrm{\Delta }x_c^{(n)}(L)x^{(n)}(y=0,L)\stackrel{~}{x}^{(n)}(y=0)=(1)^{n+1}a_ϵ^nL^{dn/\nu }\mathrm{\Delta }f_c^{(n)}$$ (29) The $`n=1`$ correction is absent by fiat: the choice of $`ϵ_c`$ ensures this. The $`n3`$ corrections are sufficiently strongly ‘irrelevant’ (they vanish sufficiently strongly with $`L`$) that they may reasonably be neglected. But the $`n=2`$ correction decays only slowly: $$\mathrm{\Delta }x_c^{(2)}(L)=a_ϵ^2L^{\alpha /\nu }\mathrm{\Delta }f_c^{(2)}=a_ϵ^2L^{\alpha /\nu }c_{0c}g(L)$$ (30) where the last step defines $`g(L)`$ (a convenient parameter) while $$c_{0c}\mathrm{\Delta }f_c^{(2)}=\frac{^2\mathrm{\Delta }f(\beta ,L)}{\beta ^2}_{\beta _c}$$ (31) is identifiable as the constant ‘background’ to the near-critical canonical specific heat. With this addition, Eq. (23) is modified to read $$L^d\left[ϵ^2_cϵ_c^2\right]L^{\alpha /\nu }a_ϵ^2x^{(2)}(y=0,L)=L^{\alpha /\nu }a_ϵ^2\stackrel{~}{x}^{(2)}(y=0)+c_{0c}$$ (32) which is now fully consistent with the MC measurements of Fig. 2, with (it is to be noted) a negative value for $`c_{0c}`$ . From a thermodynamic point of view these results simply reflect the fact that, for any system size practically accessible to computer-simulation, the ‘critical’ contribution to the canonical specific heat is not large enough to dominate the ‘non-critical background’. But the argument also shows us how to eliminate the effects of this ‘background’ from the energy pdf. Consider the cumulant representation of the scaling energy pdf (Eq. 18) at criticality: $$\stackrel{~}{P}(x|y=0)=\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑\tau e^{ix\tau }\mathrm{exp}\left[\underset{n=1}{\overset{\mathrm{}}{}}\frac{(i\tau )^n}{n!}\stackrel{~}{x}^n(y=0)\right]$$ (33) The corresponding relation for the observed energy pdf at criticality, written in its inverse form, is $$\mathrm{exp}\left[\underset{n=1}{\overset{\mathrm{}}{}}\frac{(i\tau )^n}{n!}x^{(n)}(y=0,L)\right]=_{\mathrm{}}^{\mathrm{}}𝑑x^{}e^{ix^{}\tau }P(x^{}|y=0,L)$$ (34) Appealing to the our conclusion that, for large enough $`L`$, the cumulants of the two pdfs differ significantly only in the $`n=2`$ case, and using Eqs. (29) and (30), we find that $`\stackrel{~}{P}(x|y=0)`$ $`=`$ $`{\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\tau {\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x^{}e^{i(xx^{})\tau g(L)\tau ^2/2}P(x^{}|y=0,L)`$ (35) $`=`$ $`{\displaystyle \frac{1}{\sqrt{2\pi g(L)}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x^{}e^{(xx^{})^2/[2g(L)]}P(x^{}|y=0,L)`$ (36) This result shows that the scaling form of the critical pdf may be exposed by convolution of the observed (and thus, generally, non-scaling) pdfs with gaussians whose widths are controlled by the specific heat background. Note that the argument rests on the fact that this background is negative (so that $`g(L)`$ as defined in Eq. (30) is positive). If the background constant were positive our argument would have to be restructured to prescribe the scaling form by a process of deconvolution, which is numerically problematic. As it is, the convolution process can be implemented easily. With $`c_{0c}`$ fixed by the ordinate intercept in Fig. 2, the pdfs measured on different system sizes can each be corrected in this way to yield estimates of the scaling pdf. The results are shown in Fig. 4. The improvement with respect to the raw data (Fig. 3) is striking. This improvement reflects not only the removal of the non-scaling contribution to the second cumulant but also that the requisite convolution process provides a natural smoothing of the MC data . The consequences of this correction for the shape of the distribution are also striking. The skewness clearly visible in the raw distributions (Fig. 3) is largely suppressed to expose a scaling form that is, at first appearance, gaussian. Indeed the portion of the distribution evident on the scale of Fig. 4 is gaussian to within deviations of a few per cent. However the behavior in the wings (evident on the logarithmic scale utilized in Fig. 5) is markedly different on the high- and low-energy sides. The scaling of the critical energy pdf corroborates the scaling of the microcanonical entropy (cf Eq. (18)). Given the double appearance of $`\stackrel{~}{𝒮}(x)`$ in Eq. (18) it is practical to infer only the ‘effective’ microcanonical entropy scaling function $$\stackrel{~}{𝒮}_{eff}(x)\stackrel{~}{𝒮}(x)+\frac{1}{2}\mathrm{ln}\left[\frac{\stackrel{~}{𝒮}^{\prime \prime }(x)}{2\pi }\right]$$ (37) Fig. 5 shows the form implied by Eq. (18) $$\stackrel{~}{𝒮}_{eff}(x)=\stackrel{~}{𝒮}_{eff}(x=0)+\mathrm{ln}\left[\frac{\stackrel{~}{P}(x|y=0)}{\stackrel{~}{P}(x=0|y=0)}\right]$$ (38) We note as a matter of empirical fact that $`\stackrel{~}{𝒮}_{eff}(x)`$ is concave. The concavity of $`\stackrel{~}{𝒮}(x)`$ itself is already assumed in our basic scaling ansatz . ## III The scaling theory: implications and tests Although we have no first-principles calculation of the scaling function $`\stackrel{~}{𝒮}(x)`$ to offer here we can identify, and test, some of the properties it must have to match anticipated behavior in both the thermodynamic and finite-size-critical limits. We consider, in particular, the limiting large $`x`$ behavior. In this regime we anticipate that $$\stackrel{~}{𝒮}(x)b_\pm x^\theta +r_\pm (x1)$$ (39) where the $`+`$ and $``$ subscripts refer, respectively, to the regions of positive and negative $`x`$. To make explicit identifications of the new quantities introduced in this equation (the exponent $`\theta `$ and the amplitudes $`b_\pm `$, $`r_\pm `$), consider the scaling part of the partition function (Eq. 12). In the limit of large $`y`$ the integral in Eq. (12) is dominated by one or other of the large $`x`$ regimes. Substituting Eq. (39), a saddle-point integration yields $$\mathrm{ln}\stackrel{~}{𝒵}(y)a_\pm y^{\theta /(\theta 1)}+r_\pm (y1)$$ (40) where the $`+`$ and $``$ subscripts now refer, respectively, to the regions of negative and positive $`y`$ , and $$\frac{a_+}{a_{}}=\left(\frac{b_{}}{b_+}\right)^{1/(\theta 1)}$$ (41) As in the argument leading to Eq. (4) the fluctuations about the saddle point are canceled by the pre-exponential factor in Eq. (12) to leave power-law (‘ln-free’) behavior . The thermodynamic limit of the near-critical free energy, defined by Eq. (15), now follows as $$\stackrel{~}{F}(\beta ,L)L^d\left[f_0(\beta )A_\pm \beta \beta _c^{2\alpha }\right]r_\pm $$ (42) where we have identified $$\theta =\frac{2\alpha }{1\alpha }$$ (43) and (given Eq. 41) $$\frac{A_+}{A_{}}=\frac{a_+}{a_{}}=\left(\frac{b_{}}{b_+}\right)^{1\alpha }$$ (44) To establish the role of the remaining constants ($`r_\pm `$) in Eq. (39) we consider the anomalous contribution to the free energy defined by $$F_a(\beta )\underset{L\mathrm{}}{lim}\left\{\stackrel{~}{F}(\beta ,L)L^d\underset{L\mathrm{}}{lim}\frac{\stackrel{~}{F}(\beta ,L)}{L^d}\right\}$$ (45) Appealing to Eq. (42), and recalling our sign convention , we identify $$F_a(\beta )=\{\begin{array}{cc}r_+\hfill & (\beta <\beta _c)\hfill \\ r_{}\hfill & (\beta >\beta _c)\hfill \end{array}$$ (46) On the basis of rather general arguments we expect that away from a critical point the free energy anomaly is just minus the logarithm of the number of coexisting phases, so that $`r_+`$ $`=`$ $`F_a(\beta <\beta _c)=0`$ (47a) $`r_{}`$ $`=`$ $`F_a(\beta >\beta _c)=\mathrm{ln}2`$ (47b) In the critical finite-size limit we find from Eq. (15) $$\stackrel{~}{F}(\beta _c,L)=L^df_0(\beta _c)\mathrm{ln}\stackrel{~}{𝒵}(0)$$ (48) The critical value of the free energy anomaly, defining the Privman -Fisher constant $`U_0`$ , follows as $$U_0F_a(\beta _c)=\mathrm{ln}\stackrel{~}{𝒵}(0)$$ (49) These predictions are testable to varying degrees through both the energy-dependence of the energy pdf and the temperature-dependence of the associated free energy. Figure 6 shows the results for the ratio of the specific heat amplitudes that follow from Eq. (44) when the measured decay of the critical energy pdf (Figs. 4, 5) is matched to the prediction (39), in conjunction with Eq. (18). We can expect the predictions and observations to match up only in a window of $`x`$ values. Clearly, $`x`$ must be large enough to lie within the thermodynamic critical region; but it must also not be so large that the associated energy lies outside the domain of validity of the basic scaling ansatz (Eq. 9b). The size of this window should increase with increasing system size. The location of this window on the $`x`$-axis may also be expected to be different for the positive and negative $`x`$-branches –if, as seems reasonable, one regards the correlation length $`\xi `$ (rather than $`xϵϵ_c`$ or $`y\beta \beta _c`$) as a measure of criticality. This is, indeed, the view we have adopted . Thus Fig. 6 shows the results for the ‘effective’ amplitude ratio, obtained by fitting over ranges of $`x`$ values, with each pair of (positive and negative ranges) being centered on the same value of $`z=L/\xi `$, used as the abscissa . On the basis of this data we make the assignment $`A_+/A_{}=0.575(10)`$ which is to be compared with $`A_+/A_{}=0.523(9)`$ in reference and $`0.567(16)`$ in reference . In Fig. 7 we show the results for $`\mathrm{ln}\stackrel{~}{𝒵}(y)`$ that follow (cf Eqs. (12), (37), (38)) from the measured energy pdfs, using Eq. (38). The latter determines $`\stackrel{~}{𝒮}_{eff}(x)`$ only to within an additive constant which must be fixed by appeal to the predicted value of either $`r_+`$ or $`r_{}`$. We have chosen the latter so that Eq. (47b) is satisfied, by fiat. The motivation for this choice is that it provides us with an inherently more reliable estimate of the parameter $`U_0`$ (which, unlike $`r_\pm `$, is not known a priori). Since $`U_0=\mathrm{ln}\stackrel{~}{𝒵}(0)`$ is closer to $`r_{}`$ than to $`r_+`$ the function $`\mathrm{ln}\stackrel{~}{𝒵}(y)`$ converges more quickly to its $`y>0`$ asymptote than it does to its $`y<0`$ asymptote. Fixing $`r_{}`$ (the intercept of the $`y>0`$ asymptote) thus tethers the value assigned to $`U_0`$ more effectively than fixing $`r_+`$. As with the amplitude ratio considered above, the value assigned to $`U_0`$ depends upon the range of $`y`$-values used in the fit to the anticipated asymptotic form (Eq. 40). Again we have chosen to characterize the temperature range utilized through the value of the ratio $`z=L/\xi `$; again we can expect the analysis to be trustworthy only if it is based upon data lying within the thermodynamic-critical window. Our data (Fig. 8) do not allow a systematic analysis of the approach to the desired limit; but they provide the basis for the assignment $`U_0=0.57(2)`$ . The assignment of the uncertainty limit is subjective but, we think, conservative. We note the close correspondence with the assignment ($`U_0=0.57`$) emerging from an earlier study , similar in concept, but utilizing the distribution of the order parameter. However our assignment differs (in what would seem to be a statistically significant fashion) from the result $`U_0=0.625(5)`$ obtained by Mon on the basis of altogether different techniques. ## IV Microcanonical and canonical specific heats ### A Generalities Thus far we have focused on the implications of the microcanonical entropy for observations made in the canonical ensemble. We now turn to consider their implications for observations made within ensembles that are (or are approximations to) microcanonical. We will assume (in keeping with eg ) that the temperature of a microcanonical system should be identified from the relation $$\beta ^{\mu e}(ϵ,L)=L^d\frac{\mathrm{ln}\mathrm{\Omega }(ϵ,L)}{ϵ}$$ (50) This identification is certainly required in the thermodynamic limit; but in the context of systems of finite size it is, it seems, a matter of convention . It is illuminating to link this temperature with canonical observables. Appealing to Eq. (16) we may write $$\beta ^{\mu e}(ϵ,L)=\beta +L^d\frac{\mathrm{ln}P(ϵ|\beta ,L)}{ϵ}$$ (51) where (notwithstanding appearances to the contrary) the rhs depends on $`ϵ`$ but not $`\beta `$. This result shows that the equation prescribing the microcanonical temperature for a given energy is just the inverse of the equation prescribing the most probable energy for a given temperature: $$\beta =\beta ^{\mu e}(ϵ,L)ϵ=\widehat{ϵ}^{ce}(\beta ,L)$$ (52) By comparison, within the canonical ensemble itself, ‘the’ energy for a given temperature is customarily identified with the canonical mean: $$ϵ=\overline{ϵ}^{ce}(\beta ,L)$$ (53) Eqs. (52) and (53) make it immediately plain that the energy-temperature relationships associated with the two ensembles will coincide to the extent that the canonical energy distribution is gaussian (and thus has coincident mean $`\overline{ϵ}^{ce}`$ and mode $`\widehat{ϵ}^{ce}`$). This correspondence is guaranteed in the thermodynamic limit, but not (in general) when finite-size effects are significant. The energy-temperature relationships are most usually probed through their derivatives, the associated specific heats. In the microcanonical case $$c^{\mu e}(ϵ,L)=\left[\frac{\beta ^{\mu e}(ϵ,L)}{ϵ}\right]^1=L^d\left[\frac{^2\mathrm{ln}P(ϵ|\beta ,L)}{ϵ^2}\right]^1$$ (54) where, again, the $`\beta `$-dependence of the rhs is illusory. In the canonical case (appealing to Eq. (19)) $$c^{ce}(\beta ,L)=\frac{\overline{ϵ}^{ce}(\beta ,L)}{\beta }=L^dϵ^{(2)}(\beta ,L)$$ (55) Like the two ‘caloric equations of state’ (Eqs. (52) and (53)) these two specific heats are guaranteed to agree in the thermodynamic limit; but they differ (in general) in the finite-size critical regime to which we now turn. ### B Scaling forms First, we examine the asymptotic scaling regime where the the ‘background’ contribution to the specific heat can be neglected. We will consider the consequences of the corrections associated with the latter in the following section. In the scaling regime where the canonical energy pdf may be represented by its scaling form (Eqs. (17) and (18)) Eq. (51) can be rewritten in terms of the energy and temperature scaling variables (Eqs. (7) and (13)) as $$y^{\mu e}(x)=y+\frac{\mathrm{ln}\stackrel{~}{P}(x|y)}{x}=\frac{\stackrel{~}{𝒮}_{eff}(x)}{x}$$ (56) where in the last step we have exercised the right to set $`y=0`$ (the result is independent of $`y`$) and have made use of Eqs. (18) and (37). The microcanonical specific heat (Eq. 54) follows in scaling form $$c^{\mu e}(ϵ,L)L^{\alpha /\nu }a_ϵ^2\stackrel{~}{c}^{\mu e}(x)$$ (57a) with $$\stackrel{~}{c}^{\mu e}(x)=\left[\frac{^2\mathrm{ln}\stackrel{~}{P}(x|y=0)}{x^2}\right]^1=\left[\frac{^2\stackrel{~}{𝒮}_{eff}(x)}{x^2}\right]^1$$ (57b) The scaling form of the canonical specific heat follows in a similar fashion, using Eqs. (20) and (55) $$c^{ce}(\beta ,L)L^{\alpha /\nu }a_ϵ^2\stackrel{~}{c}^{ce}(y)$$ (58a) with (Eq. 23) $$\stackrel{~}{c}^{ce}(y)=\stackrel{~}{x}^{(2)}(y)$$ (58b) The forms of both the scaling functions $`\stackrel{~}{c}^{\mu e}(x)`$ and $`\stackrel{~}{c}^{ce}(y)`$ can be determined from the scaling form for the microcanonical entropy (Fig. 5) or, equivalently, the critical canonical energy pdf (Fig. 4), established in the preceding section. They are compared in Fig. 9. In the microcanonical case we have used Eq. (56) to identify the microcanonical temperature $`y=y^{\mu e}(x)`$ to be associated with a given value of the energy variable $`x`$. In the thermodynamic limit realised at large values of $`y`$ the two functions are, necessarily, consistent with one another, and approach the asymptotic behavior implied by Eq. (42). In the finite-size-critical (small $`y`$) regime, however, clear differences between the two scaling forms are apparent. In particular, the microcanonical maximum exceeds the canonical maximum by some $`10\%`$. One can show (Appendix C) that this –the fact that the microcanonical maximum is the larger one– follows necessarily if the scaling function $`\stackrel{~}{𝒮}(x)`$ is concave. The two scaling functions cross very close to the point ($`y=0`$) identifying the bulk critical temperature. One can see this already in published microcanonical data ; a similar ‘ensemble-independence’ has also been noted in studies of the ‘gaussian ensemble’ . We have been unable to see any deep reason for this correspondence; but we do not discount the possibility that there is one. Though clearly visible, the differences between the two scaling functions are smaller than suggested by existing MC data . The next section explains why. ### C Beyond scaling: the role of the ‘background’ To understand the behavior observed in MC studies of microcanonical behavior, we must allow for the ‘corrections to scaling ’ which, in the canonical ensemble, are reflected simply in the existence of the additive negative background contribution to the heat capacity; their signature in the microcanonical ensemble is more subtle. The differences between the canonical and microcanonical results is effectively a strongly anharmonic effect: in a system of finite size, critical fluctuations sample a region of the entropy surface sufficiently large that the variation of its curvature becomes significant. We can expose the consequences analytically within an anharmonic perturbation theory in the cumulants of the energy pdf. The calculation is straightforward and we describe it in outline only. We appeal to the cumulant representation of the energy pdf at some (general) temperature: $$P(ϵ|\beta ,L)=\frac{1}{2\pi }_{\mathrm{}}^{\mathrm{}}𝑑\tau e^{iϵ\tau }\mathrm{exp}\left[\underset{n=1}{\overset{\mathrm{}}{}}\frac{(i\tau )^n}{n!}ϵ^{(n)}(\beta ,L)\right]$$ (59) We expand perturbatively to first order in the fourth cumulant and to second order in the third. We evaluate the second derivative of the logarithm of this function, which determines (cf Eq. (54)) the microcanonical specific heat associated with a given energy density. We evaluate this function at the modal energy $`\widehat{ϵ}=\widehat{ϵ}(\beta ,L)`$ associated with the chosen temperature, prescribed by the (perturbative) solution of the microcanonical caloric equation of state (Eq. 51). The result is $$c^{\mu e}(\widehat{ϵ},L)=c^{ce}(\beta ,L)\left\{1\frac{ϵ^{(4)}(\beta ,L)ϵ^{(2)}(\beta ,L)\left[ϵ^{(3)}(\beta ,L)\right]^2}{2\left[ϵ^{(2)}(\beta ,L)\right]^3}+\mathrm{}\right\}$$ (60) Eq. 19 shows that the cumulant correction terms displayed in this equation are $`O(L^d)`$ in the thermodynamic limit, confirming the equality of microcanonical and canonical predictions in this limit. To see what happens in the finite-size-critical region we focus (for simplicity) on the temperature $`\beta _m`$ for which the canonical specific heat is maximal, identified by the solution of $$ϵ^{(3)}(\beta ,L)=0=\frac{dc^{ce}(\beta ,L)}{d\beta }$$ (61) At this temperature Eq. (60) simplifies to $$c^{\mu e}(\widehat{ϵ}_m,L)=c^{ce}(\beta _m,L)\frac{L^dϵ^{(4)}(\beta _m,L)}{2ϵ^{(2)}(\beta _m,L)}+\mathrm{}$$ (62) where $`\widehat{ϵ}_m=\widehat{ϵ}(\beta _m,L)`$, and we have used Eq. (55). Now we appeal to the scaling forms for the cumulants (Eq. 20), and fold in the effects of the additional non-scaling contribution to the second cumulant (Eq. 30) to conclude that $$c^{\mu e}(\widehat{ϵ}_m,L)=c^{ce}(\beta _m,L)\frac{L^{\alpha /\nu }a_ϵ^2\stackrel{~}{x}^{(4)}(y_m)}{2\left[\stackrel{~}{x}^{(2)}(y_m)g(L)\right]}+\mathrm{}$$ (63) This result makes clear (albeit perturbatively) that, in the finite-size-limited regime, the temperature-independent additive ‘background constant’ in the canonical specific heat (manifested in the parameter $`g(L)`$) does not simply translate into an additive energy-independent background in its microcanonical counterpart. To expose the implications for the difference between canonical and microcanonical specific heats we introduce the dimensionless parameter: $$R(L)\frac{c^{\mu e}(\widehat{ϵ}_m,L)c^{ce}(\beta _m,L)}{c^{ce}(\beta _m,L)}=\frac{\stackrel{~}{x}^{(4)}(y_m)}{2\left[\stackrel{~}{x}^{(2)}(y_m)g(L)\right]^2}+\mathrm{}$$ (64) Then $$\frac{R(L)}{R(\mathrm{})}=\left[\frac{\stackrel{~}{x}^{(2)}(y_m)}{\stackrel{~}{x}^{(2)}(y_m)g(L)}\right]^2=\left[1\frac{c_{0c}}{c^{ce}(\beta _m,L)}\right]^2$$ (65) where $`R(\mathrm{})`$ is the scaling limit of $`R(L)`$. The significance of the background constant $`c_{0c}`$ –in particular, its sign– is now apparent. The negative value of this constant results in an amplification of the difference between the microcanonical and canonical results (at $`\beta _m`$), to a degree that diminishes with increasing system size. This is not simply the trivial effect that would arise from a uniform (downward) shift of both functions: Eq. (63) shows that this is not what happens, as does the power of two on the rhs of Eq. (65). It is not hard to track down the origins of this effect. The difference between the canonical and microcanonical specific heats is, we have noted, an anharmonic effect; in the present context the ‘corrections to scaling’ reduce (only) the second cumulant of the energy pdf and thus, in a relative sense, enhance the anharmonic (non-gaussian) character of the energy pdf, as one can see immediately from a comparison of Figures 3 and 4. The effect is significant. For $`L=10`$ (as used in the simulations reported in ), estimating $`c^{ce}(\beta _m,L)`$ by $`c^{ce}(\beta _c,L)`$ one can read off from Fig. 2 that $`R(L)/R(\mathrm{})4`$. The somewhat unexpected conclusion that the fractional difference between $`c^{ce}`$ and $`c^{\mu e}`$ at bulk criticality actually decreases for increasing $`L`$ is consistent with some MC studies . ## V Conclusions We review, briefly, the three principal strands of this work. First, we have broached the general question of the finite-size corrections to the density of states of a many-body system. The explicit proposal for the pre-exponential structure advanced in Eq. (2) is consistent with the pre-factor-free structure of the canonical partition function and with the behavior of the simple models discussed in Appendix B. Given the growing interest in the behavior of mesoscopically-sized systems, this proposal seems to merit some further study, with more rigor than we have attempted to offer here. Second, we have shown how one can fold out from the canonical energy fluctuation spectrum the principal corrections to scaling. The underlying behavior exhibits scale-invariance to a degree that seems remarkable, given the relative weakness of energy fluctuations. It is, we have seen, also largely consistent with established 3d-Ising critical properties. Third, we have provided a finite-size scaling theory of the microcanonical ensemble. This was the original motivation for this work– specifically, the suggestion that the finite-size-smearing of critical behavior characteristic of the canonical ensemble is ‘greatly reduced’ within the microcanonical framework. Reference offers two pieces of supporting evidence for this contention, which merit final comment. Reference suggests, firstly, that, in the vicinity of $`ϵ_c`$, the microcanonical entropy (measured with the techniques described in ) can be adequately represented by a form (Eq. 6 of reference ) which allows for no finite-size corrections at all, and which corresponds essentially to the large $`x`$ limit (Eq. 39) of our scaling function. In fact the quality of the fit provided by this representation is rather poor. And we would expect it to be so. The measured microcanonical entropy evolves in a manifestly smooth way between the limiting thermodynamic forms appropriate above and below $`ϵ_c`$; Eq. 6 of reference is non-analytic at $`ϵ_c`$. Moreover, in analyzing data for the entropy and its derivatives, it is – we have seen – essential (on all systems practically accessible) to do justice to the corrections associated with the background constant $`c_{0c}`$. Even in the thermodynamic limit the corrections allowed for in Eq. (8) of reference do not do this. The second piece of supporting evidence offered in reference is a striking enhancement of the critical peak in the microcanonical specific heat, with respect to its canonical counterpart. As we have seen, this behavior is at least partly due to the effects associated with $`c_{0c}`$; Figure 9 indicates that the underlying differences are rather less dramatic. ## ACKNOWLEDGMENTS NBW acknowledges the financial support of the Royal Society (grant no. 19076), the EPSRC (grant no. GR/L91412) and the Royal Society of Edinburgh. ## A Defining a density of states We discuss here, in general terms, the issues arising in defining a density of states function for a system in which the energy spectrum is discrete. The conventional argument makes the identification $$\mathrm{\Omega }(ϵ,L)\delta ϵ=\underset{E<E_s<E+\delta E}{}\mathrm{\Omega }_s$$ (A1) with the implicit assumption that the right hand side is proportional to $`\delta E`$ ($`L^d\delta ϵ`$). This requires that: 1. There exist many distinct levels $`s`$ within the interval $`\delta E`$. 2. The level degeneracy $`\mathrm{\Omega }_s`$ is slowly varying over the interval $`EE+\delta E`$. The fractional variation of $`\mathrm{\Omega }_s`$ over the interval $`\delta E`$ can be estimated using Eq. (50); condition C2 can then be expressed in the form $$\frac{d\mathrm{ln}\mathrm{\Omega }(ϵ,L)}{dϵ}\delta ϵ=\beta ^{\mu e}(ϵ,L)\delta E1$$ (A2) Taken together, conditions C1 and C2 thus amount to the requirement that $$\delta E_I\frac{1}{\beta ^{\mu e}(ϵ,L)}$$ (A3) where $`\delta E_I`$ characterizes the intrinsic discreteness of the energy spectrum. This condition is trivially satisfied in the classical limit (Appendix B1 considers one case explicitly). But there are obvious exceptions: in the Ising model (Appendix B2) Eq. (A3) is satisfied only at energies corresponding to microcanonical temperatures that are ‘high’ on the scale of the critical temperature. Or, to put it another way, $`\mathrm{\Omega }_s`$ is certainly not slowly varying over a range wide enough to embrace many system energy levels. We must now recognize, however, that Eq. (A1) (along with its implicit assumptions) does not faithfully reflect the conditions needed to legitimize the transition from discrete (Eq. 1a) to continuum (Eq. 1b) representations. Instead of Eq. (A1) we require, rather, that we can consistently write $$\mathrm{\Omega }(ϵ,L)e^{\beta E}\delta ϵ=\underset{E<E_s<E+\delta E}{}\mathrm{\Omega }_se^{\beta E_s}$$ (A4) where (while retaining condition C1) we must replace condition C2 by 1. $`\mathrm{\Omega }_se^{\beta E_s}`$ is slowly varying over the interval $`EE+\delta E`$, if that interval lies in a range contributing significantly to the thermal properties at temperature $`\beta `$. The range contributing significantly … is centered on the saddle point $`\widehat{E}=L^d\widehat{ϵ}`$ (Eq. 6). As a result, while condition C2 requires Eq. (A2), condition C2A requires only that $$\frac{d^2\mathrm{ln}\mathrm{\Omega }(ϵ,L)}{dϵ^2}\delta ϵ^2=L^d\left[c^{\mu e}(ϵ,L)\right]^1\delta E^21$$ (A5) where we have used Eq. (54). Thus, in place of Eq. (A3), we need simply $$\delta E_I\left[L^dc^{\mu e}(ϵ,L)\right]^{1/2}L^d\left[ϵ^{(2)}(\beta ^{\mu e},L)\right]^{1/2}$$ (A6) where the last step uses Eq. (55), and $`\beta ^{\mu e}=\beta ^{\mu e}(ϵ,L)`$ (Eq. 50). This equation expresses more explicitly the implications of condition C2A. A density of states function will exist in the operational sense (Eq. 1b) that it may be used to compute thermal properties at a given temperature as long as the canonical energy distribution (for that temperature) is broad on the scale of the intrinsic discreteness of the energy spectrum. ## B Density of states of simple models Here we show that the general form for the density of states function proposed in Eq. (2) is consistent with exact results for two simple models. ### 1 Quasi-continuous energy spectrum: harmonic lattice model Consider a system (a harmonic model of the vibrations of a crystal structure, for example) whose energy spectrum is that of $`N`$ weakly-interacting harmonic oscillators, with associated frequencies $`\nu _j,j=1\mathrm{}N`$. Then $`E(\{n\})=h{\displaystyle \underset{j=1}{\overset{N}{}}}n_j\nu _j{\displaystyle \underset{j=1}{\overset{N}{}}}ϵ_j`$ gives the energy of a microstate in which (for each $`j`$) mode $`j`$ has quantum number $`n_j`$. We consider the classical ($`h0`$) limit, in which the energy levels are quasicontinuous. In this case $`\delta E_Ih\nu _{\mathrm{m}\mathrm{i}\mathrm{n}}`$, Eq. (A3) is satisfied, and we may proceed as in Eq. (A1) to write $`\mathrm{\Omega }(ϵ,N)={\displaystyle \frac{1}{\delta ϵ}}{\displaystyle \underset{\{n\}}{}}D(E(\{n\})Nϵ)`$ where $`D(X)=\{\begin{array}{cc}1\hfill & \text{if }0<X<N\delta ϵ\hfill \\ 0\hfill & \text{otherwise}\hfill \end{array}`$ while $`ϵE/N`$ is the energy per oscillator. In the $`h0`$ limit the sums on $`n_j`$ can be replaced by integrals on $`ϵ_j`$ to give $`\mathrm{\Omega }(ϵ,N)=ϵ^1E^NQ(N)I(N)`$ where $`I(N)={\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle _0^1}𝑑x_j\delta (1{\displaystyle \underset{j}{}}x_j)`$ and $`Q(N)={\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle \frac{1}{h\nu _j}}`$ Writing an integral representation of the $`\delta `$-function we find $`I(N)={\displaystyle \frac{1}{2\pi }}{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑he^{ih}\left[{\displaystyle \frac{e^{ih}1}{ih}}\right]^N={\displaystyle \frac{2}{\pi }}{\displaystyle _0^{\mathrm{}}}𝑑hcos\left[(N2)h\right]\left[{\displaystyle \frac{\mathrm{sin}h}{h}}\right]^N={\displaystyle \frac{1}{\mathrm{\Gamma }(N)}}`$ which may be approximated using the asymptotic expansion for the $`\mathrm{\Gamma }`$ function $$\mathrm{\Gamma }(z)=\sqrt{2\pi }z^{z\frac{1}{2}}e^z\left[1+O(z^1)\right]$$ (B1) In analyzing the remaining (energy-independent) contribution we suppose that the frequency spectrum is that of a $`d=1`$ system of particles, with a gap. Then $`\mathrm{ln}Q(N)=N\mathrm{ln}hNq_N`$ where the sum $`q_N`$ may be written in the form $`q_N={\displaystyle \frac{1}{N}}{\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{ln}\nu _j={\displaystyle \frac{1}{N}}{\displaystyle \underset{r=0}{\overset{N1}{}}}H({\displaystyle \frac{2\pi r}{N}})`$ where $`H(\theta )`$ is periodic, and (invoking the assumed gap) $`H(0)`$ is non-zero. It can be shown that the $`N\mathrm{}`$ limit of this sum $`q_{\mathrm{}}={\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}H(\theta )𝑑\theta `$ has finite-size corrections that are exponentially small in N. Gathering these results together we conclude that the density of states has the form of Eq. (2) with the identifications $`N=L^d`$ and $$s(ϵ)=\mathrm{ln}\left[\frac{ϵ}{h}\right]q_{\mathrm{}}+1$$ (B2) Appealing to Eq. (4) one can readily recover, as a check, the canonical partition function $`Z(\beta ,N)=(\beta h)^Ne^{Nq_{\mathrm{}}}`$ ### 2 Discrete energy spectrum: 1d Ising model Consider a $`d=1`$ Ising model of $`N`$ sites, with periodic boundary conditions. Choosing the ground state as the energy-zero, the energy density for a macrostate of $`M`$ domain walls is $`ϵ=Mϵ_I/N`$, where $`ϵ_I`$ is the domain wall energy. The number of microstates corresponding to macrostate $`M,N`$ is $`\mathrm{\Omega }_{M,N}={\displaystyle \frac{2\times N!}{(NM)!M!}}`$ Appealing to the asymptotic form (B1) once more we find that $`\mathrm{\Omega }_{M,N}=2\sqrt{{\displaystyle \frac{N}{2\pi }}}\left[x(1x)\right]^{1/2}x^{Nx}(1x)^{N(1x)}\left[1+O(N^1)\right]`$ where $`xM/N=ϵ/ϵ_I`$. In this case $`\delta E_I=ϵ_I`$ and Eq. (A3) is not in general satisfied. But, since Eq. (A6) is, we may still identify a density of states by $`\mathrm{\Omega }(ϵ,N)={\displaystyle \frac{1}{2ϵ_I}}\mathrm{\Omega }_{M,N}`$ which one may then readily recast in the form of Eq. (2) with the identifications $`N=L^d`$ and $$s(ϵ)=(1x)\mathrm{ln}(1x)x\mathrm{ln}x$$ (B3) Again, as a check, one can readily use this result to recover the canonical free energy density in the form $`f(\beta ,N)={\displaystyle \frac{1}{N}}\mathrm{ln}Z(\beta ,N)=\mathrm{ln}(2\mathrm{cosh}K)+K`$ where $`2K=\beta ϵ_I`$, and the last term reflects our choice of ground state energy. ## C A bound on the canonical specific heat We outline here an argument establishing that the maximum of the value of the microcanonical specific heat provides an upper bound for the canonical specific heat, within the asymptotic scaling region. The argument assumes the concavity of the function $`\stackrel{~}{𝒮}_{eff}(x)`$; the concavity of $`\stackrel{~}{𝒮}(x)`$ is already presupposed in the formulation of Eq. (2). We write the scaling function for the energy pdf (Eq. 18) in the form: $$\stackrel{~}{P}(x|y)=Q(x,y)G(x\widehat{x}(y),1/\stackrel{~}{S}_{eff}^{\prime \prime }(x^{}))$$ (C1) where $`G(z,b)`$ is a gaussian of zero mean, and variance $`b`$; $`\widehat{x}(y)`$ is the modal scaled energy, for a given $`y`$, the solution of $$\frac{d\stackrel{~}{𝒮}_{eff}(x)}{dx}=y$$ (C2) and $`x^{}`$ locates the maximum of the microcanonical specific heat, identified by the condition (Eq. 57b) $$\stackrel{~}{S}_{eff}^{\prime \prime }(x)\stackrel{~}{S}_{eff}^{\prime \prime }(x^{})$$ (C3) The function $`Q(x,y)`$ introduced in Eq. (C1) is defined by: $$Q(x,y)=Q_0e^{T(x,y)}$$ (C4) where $$T(x,y)=xy+\stackrel{~}{𝒮}_{eff}(x)\frac{\stackrel{~}{S}_{eff}^{\prime \prime }(x^{})}{2}(x\widehat{x}(y))^2$$ (C4) while $`Q_0`$ is an $`x`$-independent constant, defined by normalization. From the assumed concavity of $`\stackrel{~}{𝒮}_{eff}(x)`$ it is straightforward to show that, for any given $`y`$, $`T(x,y)`$ is concave in $`x`$, with a single maximum at $`x=\widehat{x}(y)`$. Now appealing to Eqs. (C1) and (58b) we can write $`\stackrel{~}{c}^{ce}(y)`$ $`=`$ $`\stackrel{~}{x}^{(2)}(y)`$ (C5) $`=`$ $`{\displaystyle 𝑑x\stackrel{~}{P}(x|y)\left[x\stackrel{~}{x}^{(1)}(y)\right]^2}`$ (C6) $``$ $`{\displaystyle 𝑑x\stackrel{~}{P}(x|y)\left[x\widehat{x}(y)\right]^2}`$ (C7) $`=`$ $`{\displaystyle 𝑑zQ(\widehat{x}(y)+z,y)G(z,1/\stackrel{~}{S}_{eff}^{\prime \prime }(x^{}))z^2}`$ (C8) $`=`$ $`{\displaystyle 𝑑z\stackrel{~}{Q}(z,y)G(z,1/\stackrel{~}{S}_{eff}^{\prime \prime }(x^{}))z^2}`$ (C9) where $`\stackrel{~}{Q}(z,y)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left[Q(\widehat{x}(y)+z,y)+Q(\widehat{x}(y)z,y)\right]`$ (C10) $`=`$ $`{\displaystyle \frac{Q_0}{2}}\left[e^{T(\widehat{x}(y)+z,y)}+e^{T(\widehat{x}(y)z,y)}\right]`$ (C11) From the properties of the function $`T(x,y)`$ it is straightforward to show that $`\stackrel{~}{Q}(z,y)`$ has a single turning point (at $`z=0`$), and that there exists some $`z_0(y)`$ such that $$\stackrel{~}{Q}(z,y)\{\begin{array}{cc}>1\hfill & \text{ if }z<z_0(y)\hfill \\ <1\hfill & \text{ if }z>z_0(y)\hfill \end{array}$$ (C12) Then, finally, appealing to Eqs. (C9) and (57b) $`\stackrel{~}{c}^{ce}(y)\stackrel{~}{c}^{\mu e}(x^{})`$ $`<`$ $`{\displaystyle 𝑑z\left[\stackrel{~}{Q}(z,y)1\right]G(z,1/\stackrel{~}{S}_{eff}^{\prime \prime }(x^{}))z^2}`$ (C13) $`<`$ $`z_0^2(y){\displaystyle 𝑑z\left[\stackrel{~}{Q}(z,y)1\right]G(z,1/\stackrel{~}{S}_{eff}^{\prime \prime }(x^{}))}`$ (C14) $`=`$ $`0`$ (C15) where the last step exploits normalization conditions. It follows that the microcanonical specific heat maximum $`\stackrel{~}{c}^{\mu e}(x^{})`$ provides an upper bound for the canonical specific heat.
no-problem/9908/astro-ph9908276.html
ar5iv
text
# REFERENCES Over the last few years a number of planets orbiting nearby stars have been discovered (for a review and references see). Their existence has been inferred from tiny doppler shifts in the light from the star due to its orbit around the center of mass. The periodicity and magnitude of the doppler shifts can be used to determine the mass and orbit of the planet. Surprisingly about half a dozen large planets (ie. with Jupiter sized masses or larger) have been found which have orbits very close to the star (between $`0.04`$ to $`0.06`$ AU). We summarize the data in Table 1 which lists all of the confirmed planets with orbits less than $`0.1`$ AU. | STAR | MASS (J) | SEM-MAJ AXIS (AU) | PERIOD (d) | ECC | | --- | --- | --- | --- | --- | | HD75289 | 0.42 | 0.046 | 3.51 | 0.054 | | 28.94 pc | | | | | | 51 Peg | 0.47 | 0.05 | 4.23 | 0.0 | | 15.36 pc | | | | | | HD187123 | 0.52 | 0.042 | 3.097 | 0.03 | | 49.92 pc | | | | | | Ups And | 0.71 | 0.059 | 4.617 | 0.034 | | 13.47 pc | | | | | | HD217107 | 1.28 | 0.04 | 7.11 | 0.14 | | 19.72 pc | | | | | | Tau Boo | 3.87 | 0.0462 | 3.3128 | 0.018 | | 15.60 pc | | | | | | HD98230 | 37 | $``$ 0.06 | 3.98 | 0.00 | | HD283750 | 50 | $``$ 0.04 | 1.79 | 0.02 | Table Caption: Table obtained from Ref. showing the mass (in units of Jupiter mass), Semi-Major axis, period (in days) and eccentricity of all of the confirmed planets with orbits less than $`0.1`$ AU. Whilst the doppler shift detection technique is quite sensitive to large mass planets with close orbits, the fact that any such planets with these properties exist at all is unexpected. To date none of these nearby planets has been seen directly, however this should be possible in the future provided that they are made of ordinary matter. We believe that an interesting alternative possibility exists and that is that these close orbiting large planets might be made of mirror matter. \[This has also been suggested independently by Volkas. \]. The existence of mirror matter is well motivated from a particle physics point of view, since these particles are predicted to exist if parity and indeed time reversal are unbroken symmetries of nature. The idea is that for each ordinary particle, such as the photon, electron, proton and neutron, there is a corresponding mirror particle, of exactly the same mass as the ordinary particle. For example, the mirror proton and the ordinary proton have exactly the same mass The mass degeneracy of ordinary and mirror matter is only valid provided that the parity symmetry is unbroken, which is the simplest and theoretically most attractive possibility. For some other possibilities, which invoke a mirror sector where parity is broken spontaneously (rather than being unbroken), see Ref... Furthermore the mirror proton is stable for the same reason that the ordinary proton is stable, and that is, the interactions of the mirror particles conserve a mirror baryon number. The mirror particles are not produced in Laboratory experiments just because they couple very weakly to the ordinary particles. In the modern language of gauge theories, the mirror particles are all singlets under the standard $`GSU(3)SU(2)_LU(1)_Y`$ gauge interactions. Instead the mirror particles interact with a set of mirror gauge particles, so that the gauge symmetry of the theory is doubled, i.e. $`GG`$ (the ordinary particles are, of course, singlets under the mirror gauge symmetry). Parity is conserved because the mirror particles experience $`V+A`$ mirror weak interactions and the ordinary particles experience the usual $`VA`$ weak interactions. Ordinary and mirror particles interact with each other predominately by gravity only. At the present time there is a range of experimental evidence supporting the existence of mirror matter. Firstly, it provides a natural candidate for dark matter, which might be mirror stars (and mirror dust, planets etc). There is an interesting possibility that these mirror stars have already been detected experimentally in the MACHO experiments. Secondly, ordinary and mirror neutrinos are maximally mixed with each other if neutrinos have mass. This nicely explains the solar and atmospheric neutrino anomalies. The idea is also compatible with the LSND experiment. Interestingly, maximal ordinary - mirror neutrino oscillations do not pose any problems for big bang nucleosynthesis (BBN) and can even fit the inferred primordial abundances better than the standard model. Of course due to the nature of mirror matter, its existence is difficult to rigorously prove (or disprove). If many nearby stars have close orbiting mirror planets then this should help establish the existence of mirror matter. Mirror planets cannot be seen directly because they cannot reflect the light from the star. This is a definite and in fact testable prediction of our mirror planet hypothesis. Another implication of the mirror planet hypothesis is that it may not be possible to detect the change of brightness when the planet occults the star (i.e. when the planet passes between the star and its line of sight as seen from the earth). This is just because mirror matter may be completely transparent to ordinary light! Actually it is possible that there is a small electromagnetic coupling between ordinary and mirror matter arising from photon - mirror photon kinetic mixing (see latter discussion). This small interaction can potentially make the mirror planet opaque (this will obviously depend on the strength of this interaction, i.e. the parameter $`\zeta `$ in Eq.(1) below, on the frequency of light and the amount and chemical composition of the mirror planet). Also, it should be noted that the mirror planet might be only partially opaque leading to absorption lines. This could provide a clear signal that the planet is made of mirror matter. Amusingly, if this could be observed, then it might be possible to determine the chemical composition of the mirror planet.. Another implication of the mirror planet hypothesis is that the orbital plane of the mirror planets may be in a completely different plane to ordinary planets. For example, the star Upsilon Andromedae has three confirmed Jovian planets. One with a close orbit (listed in the fourth row of the table) and two more distant planets with orbits 0.83 AU and 2.5 AU. If the two distant planets are made of ordinary matter than it is likely that the these two planets should orbit in a different plane to the close orbit planet if this is a mirror planet. This is just because ordinary and mirror matter interacts with each other predominately by gravity only. If these close orbiting planets are made of mirror matter, then a number of questions naturally arise. Firstly there are arguments which suggest that ordinary and mirror matter should be segregated on relatively large scales. A scale of $`10^5`$ stars was estimated in Ref.. This was assuming that ordinary and mirror matter interact only gravitationally (along with a number of other assumptions which may not always be valid). However, it was pointed out in Ref. that ordinary and mirror matter can interact weakly due to photon kinetic mixing. In field theory this is described by the interaction Lagrangian density $$=\zeta F^{\mu \nu }F_{\mu \nu }^{},$$ (1) where $`F^{\mu \nu }`$ ($`F_{\mu \nu }^{}`$) is the field strength tensor for electromagnetism (mirror electromagnetism). This type of Lagrangian term is gauge invariant and renormalizable and can exist at tree level or maybe induced radiatively in models without $`U(1)`$ gauge symmetries (such as grand unified theories). The effect of ordinary photon - mirror photon kinetic mixing is to give the mirror charged particles a small electric charge. That is, they couple to ordinary photons with charge $`\zeta e`$. This small non-gravitational force will allow some ordinary matter \- mirror matter collisions which can dissipate energy and help ordinary stars attract a significant (‘significant’ means of the order of 0.1 percent by mass) amount of mirror matter during their formation. Of course one may wonder why some stars have mirror companions and not other stars. In particular their is obviously no such large mirror planet orbiting our sun. One possibility is that all stars in our region of the galaxy attracted a significant amount of mirror matter during their formation. However for some of these stars (including our sun) it might be that the mirror matter was so close to the ordinary matter that it was either destroyed by tidal forces or the tidal forces prevented it from forming in the first place. Indeed, a mirror (or ordinary) planet would be destroyed by tidal forces when its radius is within the Roche limit, given by: $$r<f_R\left(\frac{\overline{\rho }_s}{\overline{\rho }_p}\right)^{1/3}R_s,$$ (2) where $`\overline{\rho }_s,\overline{\rho }_p`$ are the average densities of the star and planet respectively and $`R_s`$ is the radius of the star. Also $`f_R2.5`$ . If the mirror planet is close enough to break apart by the tidal forces then one may be left with rings of mirror matter surrounding the star. This might be similar to the rings of saturn which may have formed from the tidal break up of a moon or moons. Naturally we would expect much of the mirror ring material to have migrated to the center of the sun by either gravitational dissipation, collisions of the orbiting mirror particles with themselves and/or through the small possible non-gravitational force arizing form photon-mirror photon kinetic mixing. The latter effect can easily be estimated and any mirror matter within the radius of the star would have migrated to the center provided that $`\zeta \stackrel{>}{}10^{15}`$. This is consistent with the experimental and BBN bounds on $`\zeta `$ which imply that $`\zeta \stackrel{<}{}10^610^8`$. In summary, we have suggested that the close orbiting planets discovered in nearby stars can be plausibly explained in terms of mirror matter. Stars (such as our sun) without close orbit mirror planets may also have a significant amount of mirror matter which has broken up under the tidal forces. In this case, most of this mirror matter would be expected to have migrated to the center of the sun, although some mirror material may also exist in the form of mirror rings (similar to the rings of saturn, except they are made of mirror matter). This mirror planet hypothesis will be testable in future experiments. Acknowledgement The author thanks R. R. Volkas for his comments on the paper and A. Ignative, G. Joshi, B. Morgan and M. Drinkwater for discussions. The author also thanks C. Feynman and Z. Silagadze for constructive correspondence which have lead to significant improvements to the paper. The author is an Australian Research Fellow.
no-problem/9908/hep-ph9908495.html
ar5iv
text
# Two Meson Scattering Amplitudes and their Resonances from Chiral Symmetry and the N/D MethodTalk given at PANIC99, Uppsala (Sweden), June 10-16, 1999. ## 1 Introduction $`\chi PT`$ can be supplied with the exchange of explicit resonance fields . In doing this, a resummation up to an infinite order in the chiral expansion can be achieved from the expansion of the bare propagator of a resonance. In fact, at $`𝒪(p^4)`$, it is seen that the $`L_i`$ counterterms of $`\chi PT`$ are saturated by the exchange of the resonances. However, the amplitudes that can be built directly from $`\chi PT`$ at $`𝒪(p^4)`$ plus resonance exchanges as in , need a unitarization procedure in order to compare directly with experimental data (phase shifts, inelasticities…) for the different energy regions, in particular, around the resonance masses. This is one of the aims of the present study. On the other hand, it is well known that the scalar sector is much more controversial than the vector or tensor ones. In the latter case, the associated spectroscopy can be understood in terms of first principles coming directly from QCD as Chiral Symmetry and Large $`N_c`$ plus unitarity, once we admit VMD as dictated by phenomenlogy. We try to make use in this work of the same principles than before, that is, Chiral Symmetry, Large $`N_c`$ and unitarity in coupled channels, in order to study the scalar resonant channels. ## 2 Formalism We consider the influence of the unphysical cuts perturbatively. In this we take the zero order approach, that is, we neglect it. In we make estimations of the unphysical cuts contribution and find them to be only a few per cent of our final amplitudes. In it is discussed how to include them up to one loop calculated at $`𝒪(p^4)`$. When taking into account the N/D method the most general structure that an elastic partial wave amplitude, $`\text{T}_L`$, has when the unphysical cuts are neglected, is : $`\text{T}_L^{}(s)`$ $`=`$ $`{\displaystyle \frac{1}{\text{D}_L(s)}}`$ $`\text{D}_L(s)`$ $`=`$ $`{\displaystyle \frac{(ss_0)^{L+1}}{\pi }}{\displaystyle _{s_{th}}^{\mathrm{}}}𝑑s^{}{\displaystyle \frac{\nu (s^{})^L\rho (s^{})}{(s^{}s)(s^{}s_0)^{L+1}}}+{\displaystyle \underset{m=0}{\overset{L}{}}}a_ms^m+{\displaystyle \underset{i}{\overset{M_L}{}}}{\displaystyle \frac{R_i}{ss_i}}`$ (1) where $`\text{T}_L^{}(s)=\text{T}_L(s)/\nu ^L`$, $`\nu ^L=p^{2L}`$ with $`p`$ the center mass three momentum of the particles and $`s_0`$ is the subtraction point. In Large $`N_c`$ counting rules $`\text{T}_L1/N_c`$ and hence $`\text{D}_LN_c`$. We then split the $`\text{D}_L`$ in two parts: $`\text{D}_L^{\mathrm{}}`$ $`=`$ $`{\displaystyle \underset{m=0}{\overset{L}{}}}a_m^Ls^m+{\displaystyle \underset{i}{\overset{M_L}{}}}{\displaystyle \frac{R_i}{ss_i}}𝒪(N_c)`$ $`g(s)`$ $`=`$ $`{\displaystyle \frac{(ss_0)^{L+1}}{\pi }}{\displaystyle _{s_{th}}^{\mathrm{}}}𝑑s^{}{\displaystyle \frac{\nu (s^{})^L\rho (s^{})}{(s^{}s)(s^{}s_0)^{L+1}}}+{\displaystyle \underset{m=0}{\overset{L}{}}}a_m^{SL}s^m𝒪(1)`$ (2) where $`a_m^{L(SL)}`$ is the $`𝒪(N_c)`$($`𝒪(1)`$) of the coefficient $`a_m`$. In the large $`N_c`$ limit a partial wave amplitude is given by local plus pole terms. The local terms are taken from the lowest order $`\chi PT`$ amplitudes and the resonant ones from the exchange of resonances as given in . In doing this, we are assuming the concept of Meson Resonance Dominance, in the sense that all the local terms of order higher than two in the $`\chi PT`$ expansion are saturated by the resonance exchanges. This was shown to be case at $`𝒪(p^4)`$ in . In this way $$\text{T}_L^{\mathrm{}}=\frac{1}{\text{D}_L^{\mathrm{}}}=T^{(2)}+T^R$$ (3) where $`T^{(2)}`$ refers to the lowest order $`\chi PT`$ amplitude and $`T^R`$ to the resonance exchange amplitudes. In it is proved that $`\text{D}_L^{\mathrm{}}`$ has enough room to accommodate $`T^{(2)}+T^R`$ as given above. Thus, our final formula will be: $$\text{T}_L=p^L\left(\left[T^{(2)}+T^R\right]^1+g(s)\right)^1p^L$$ (4) The extension of the former formula to the coupled channel case is straightforward in a matrix formalism. Then, we have: $`T^{(2)}`$, $`T^R`$ $``$ $`T_{ij}^{(2)}`$, $`T_{ij}^R`$ and $`g(s)`$ and $`p`$ are now diagonal matrices . ## 3 Vectors Making use of the former formalism we study simultaneously the $`\rho `$ and $`K^{}`$ resonances . We obtain the expected leading behaviours of their bare poles and the KSFR value of the coupling with a relative deviation of only a $`6\%`$. The former conclusions are obtained after fitting the elastic P-wave $`\pi \pi `$ and $`K\pi `$ phase shifts in terms of two parameters: the subtraction constant $`a^{SL}`$ present in $`g(s)`$ and the coupling of the vector octet of resonances. ## 4 Scalars We study the $`SU(3)`$ connected meson-meson S-waves with $`I=0`$, 1 and 2, where $`I`$ refers to the isospin. Contrary with the former vector channels, which are essentially elastic, for the proposed scalar channels one has to take into account the effect of coupling channels. We include the following channels: $$\begin{array}{cc}I=0\hfill & \pi \pi (1),K\overline{K}(2),\eta \eta (3)\hfill \\ I=1\hfill & \pi \eta (1),K\overline{K}(2)\hfill \\ I=1/2\hfill & K\pi (1),K\eta (2)\hfill \end{array}$$ (5) These are the most relevant channels up to $`\sqrt{s}1.2`$ GeV. For energies higher than these other channels become increasingly more important as four pions for $`I=0`$ or $`K\eta ^{}`$ $`I=1/2`$. Thus, a proper study of the set of resonances that we find around 1.4 GeV would require the inclusion of those relevant channels . We fit, up to $`\sqrt{s}1.4`$ GeV, the following data: $`I=0`$ elastic $`\pi \pi `$ phase shifts, $`I=0`$ $`K\overline{K}\pi \pi `$ phase shifts, $`\frac{1\eta _{00}^2}{4}`$ where $`\eta _{00}`$ is the inelasticity with $`I=0`$, $`I=1/2`$ elastic $`K\pi `$ phase shifts and a mass distribution for $`I=1`$ around the $`a_0(980)`$ resonance. In a first glance at the PDG one can think that there should be at least two scalar nonets, one with a mass below 1 GeV and another around 1.4 GeV. We first include two scalar nonets but then the fit gives a remarkable feature. The octet around 1 GeV has vanishing coupling constants and the same occurs with the singlet around 1.4 GeV. As a consequence, one can reproduce the scalar data with only a singlet around 1 GeV and an octet around 1.4 GeV. The $`\chi ^2`$ per degree of freedom obtained is almost 1 with 188 experimental points. From the point of view of the resonance content of our amplitudes, the explicitly included octet with a mass about 1.4 GeV evolves to give rise to poles with masses very close to those of the $`K_0^{}(1430)`$, $`a_0(1450)`$ and $`f_0(1500)`$ resonances. In turn, the singlet around 1 GeV evolves to the physical pole of the $`f_0(980)`$. However, together with the former poles we also find other ones which do not originate from any preexisting resonance ($`T^R=0`$). They are meson-meson resonances originating from the unitarization of the lowest order $`\chi PT`$ amplitudes. On the other hand, since loops are suppressed in large $`N_c`$ these resonances disappear for $`N_c\mathrm{}`$. They correspond to the $`\sigma (500)`$, $`a_0(980)`$, $`\kappa `$ and a strong contribution to the physical $`f_0(980)`$ resonance. Thus, the $`f_0(980)`$ resonance results from two effects: a preexisting resonance around 1 GeV and a strong $`K\overline{K}`$ threshold effect. ## 5 Conclusions We have presented a systematic procedure to unitarize the tree level amplitudes coming from lowest order $`\chi PT`$ and the explicit exchange of resonance fields. We have used this method to study the vector and scalar resonances. For the vectors, we reproduce the well known features of VMD and the KSFR value for the coupling of the vector resonances $`\rho `$ and $`K^{}`$. For the controversial scalar channel, the situation is more complicated. After reproducing a large amount of experimental data, we have observed two sets of resonances. Those resonances preexisting to the unitarization: one octet around 1.4 GeV and a singlet around 1 GeV that evolves to the physical $`f_0(980)`$ resonance. The other set corresponds to meson-meson resonances with a mass $``$1 GeV: $`\sigma `$, $`a_0(980)`$, $`\kappa `$ and a strong contribution to the $`f_0(980)`$ from the $`K\overline{K}`$ threshold. This set of resonances forms a nonet and in fact when we go in our formalism to the $`SU(3)`$ limit they form an octet of degenerate resonances plus a singlet.
no-problem/9908/hep-lat9908031.html
ar5iv
text
# Bound on the string tension by the excitation probability for a vortex ## Abstract A lower bound on the string tension for large beta in SU(2) LGT is derived. The derivation is from first principles and bounds the string tension from below by the expectation for the excitation of a single ‘tagged’ thick vortex winding around the lattice. Thus confinement follows if this expectation remains nonvanishing at large beta. Numerical simulations are presented to show that this is indeed the case. Over the last two years the center vortex picture of confinement has undergone substantial development by a series of numerical investigations as well as new analytical results (see e.g. and references therein, and contributions to these proceedings). Here we report on a new analytical result relating the existence of nonzero string tension to the excitation probability of a vortex at weak coupling. The relation implies that nonvanishing expectation for an (arbitrarily) long vortex is a sufficient condition for confinement at weak coupling. We then present a measurement of this probability by numerical simulations. We consider the Wilson loop $`W[C]=\mathrm{tr}U[C]`$ in the $`SU(2)`$ LGT with plaquette action $$A_p(U)=\beta _A|\mathrm{tr}U_p|+\beta \mathrm{tr}U_p,$$ (1) and for large $`\beta `$, $`\beta _A`$. Our result is the following bound on the Wilson loop: $$W[C]\mathrm{exp}(\rho (\beta )|A|)$$ (2) where $`|A|`$ is the minimal loop area, and $`\rho (\beta )`$ $`=`$ $`(\text{Const})\mathrm{ln}[\mathrm{\hspace{0.17em}1}+\theta _^{}p^{(+)}\mathrm{tanh}K_0]`$ (3) $``$ $`(\text{Const})e^{4\beta }\theta _^{}p^{(+)}.`$ with $`K_0\frac{1}{2}\mathrm{ln}\mathrm{coth}(2\beta )`$. The quantity $`\theta _^{}p^{(+)}`$ denotes the expectation of an operator creating magnetic flux forced to wind completely around the (periodic) lattice. It is defined as follows. Define for a $`3`$-cube $`c`$: $$\eta _c=\underset{pc}{}\text{sign}\mathrm{tr}U_p.$$ (4) A configuration such that $`\eta _c=1`$ represents a Dirac monopole of $`Z(2)`$ flux residing on the cube $`c`$. The set of cubes on which $`\eta _c=1`$ must form coclosed sets, i.e. closed sets on the dual lattice (Bianchi identity - magnetic current conservation). Thus in $`d=4`$ (where a cube is dual to a bond), a coclosed set of cubes is a closed loop of dual bonds: a $`Z(2)`$ monopole loop. The smallest such set is the coboundary of a plaquette (figure 1). By virtue of the conservation constraint, the presence (absence) of a minimal length monopole ‘loop’ on the coboundary $`^{}p`$ of a plaquette $`p`$ is characterized by $`\theta _^{}p=1`$ ($`\theta _^{}p=0`$), where $$\theta _^{}p\underset{c^{}p}{}\frac{1}{2}[\mathrm{\hspace{0.17em}1}\eta _c].$$ (5) It is the expectation $`\theta _^{}p^{(+)}`$ of this quantity (5), for some fixed plaquette $`p`$, that occurs in (3). The choice of $`p`$ is irrelevant by translational invariance. The ‘$`+`$’ in $`^{(+)}`$ signifies that the expectation of $`\theta _^{}p`$, is computed with: 1. plaquette action: $`(\beta _A+\beta )|\mathrm{tr}U_p|`$; 2. $`\eta _c=1`$ for all $`3`$-cubes on the lattice other than the cubes belonging to $`^{}p`$ in the numerator, and for all $`3`$-cubes in the denominator (i.e. the partition function) of the expectation; 3. $`\mathrm{sign}U_p^{}=1`$ for every plaquette $`p^{}`$ on the $`2`$-dimensional plane $`S`$ spanning the lattice and containing the plaquette $`p`$ of $`^{}p`$. It is important to note, and easily seen that, with periodic boundary conditions, (i)-(iii) imply that $`\theta _^{}p^{(+)}`$ depends only on coset $`SU(2)/Z(2)SO(3)`$ rather than $`SU(2)`$ bond variable configurations. The expectation $`\theta _^{}p^{(+)}`$ is now seen to have the following physical meaning. (We refer to , for a physical discussion of vortices and constraints on fluxes on the lattice.) Constraint (iii) forbids any net vortex flux from crossing the surface $`S`$. Furthermore, monopole excitation is forbidden everywhere by constraint (ii) except on the coboundary of the plaquette $`p`$ on $`S`$, where the factor $`\theta _^{}p`$ enforces the presence of a monopole loop. The Dirac sheet in the $`SO(3)`$ configuration, representing the vortex flux attached to this monopole loop, is thus forced to wind around the periodic lattice in the perpendicular directions. $`\theta _^{}p^{(+)}`$ is then the excitation probability amplitude for a vortex completely winding around the lattice in the directions perpendicular to the surface $`S`$ ‘anchored’ by a minimal-length monopole loop, encircing $`S`$ (figure 2). This is in fact a version of a ’t Hooft loop operator . Shrinking the monopole loop to a point, would result in the flux of a complete, ‘unpunctured’ vortex trapped inside the periodic lattice. Eqs. (2)-(3) then imply nonvanishing string tension provided the expectation $`\theta _^{}p^{(+)}`$ remains nonvanishing in the large volume limit at (arbitrarily) large $`\beta `$. What decides this is whether the free-energy cost of the flux forced to wind around the (periodic) lattice remains finite in the large lattice limit. This can happen if flux can efficiently spread out in the transverse directions to compensate for the cost along the winding directions , . The monopole loop itself represents a local effect of fixed action cost, and serves to ‘anchor’ and tag the vortex. This is a very convenient device that leads to derivation of (2)-(3). Indeed, the derivation relies on a factorization inequality for the expectation of $`n`$ tagged vortices in terms of the product of the expectation of $`(n1)`$ tagged vortices times that of a single tagged vortex. It further uses the $`SO(3)\times Z(2)`$ formulation of the $`SU(2)`$ LGT , and subsequent duality transformation on the $`Z(2)`$ part. The somewhat lengthy details will appear elsewhere. The finite local cost associated with the monopole loop site can in fact be explicitly extracted, and then $`\theta _^{}p^{(+)}`$ can be related to the expectation for a single unpunctured vortex winding around the lattice, i.e. essentially ’t Hooft’s magnetic-flux free-energy order parameter . Now both the latter and $`\theta _^{}p^{(+)}`$ are nonvanishing in the large volume limit already when evaluated in the semiclassical approximation for spacetime dimension $`d4`$. One expects this to persist and in fact be improved in the full theory where flux can spread out nonperturbatively. In the absence of an analytical proof, we have resorted to numerical evaluation of the vortex excitation expectation (magnetic flux free-energy). Measurement was performed by the multihistogram method . The method tends to be computationally expensive. It was used in to compute the free energy of a Z(2) monopole pair as a function of the pair’s separation. The result of our computation is shown in figure 3. The lattice spacings are $`a=0.119`$ fm and $`a=0.085`$ fm for $`\beta =2.4`$ and $`\beta =2.5`$, respectively. As expected by physical reasoning, not only does the vortex free energy cost remain finite as the lattice volume grows, but it tends to zero, i.e. the weighted probability for the presence of a vortex goes to unity for sufficiently large lattice. This reflects the exponential spreading of color-magnetic flux in a confining phase. In conclusion, we have seen that nonvanishing expectation for an arbitrarily long spread-out vortex is sufficient to ensure that the $`SU(2)`$ LGT remains in a confining phase at any large $`\beta `$. This should be combined with the result that the presence of vortices at weak coupling is also necessary for confinement: constraining a Wilson loop to be insensitive to the presence of thick vortices linking with it leads to nonconfining behavior . Both these statements are clearly demonstrated in the numerical simulations. In fact, the measurement of the magnetic-flux free-energy presented here is interesting in its own right as it also shows that the probability at large $`\beta `$ for the presence in the vacuum of a sufficiently spread-out vortex actually tends to one. This is in complete agreement with the results of the closely connected measurements in . This work was supported by FOM, and NSF grant NSF-PHY 9531023.
no-problem/9908/quant-ph9908042.html
ar5iv
text
# On the Way to Submicroscopic Description of Nature11footnote 1Indian Journal of Theoretical Physics 49, no. 2, pp. 81-95 (2001); also arXiv.org e-print archive http://arXiv.org/abs/quant-ph/9908042 ## 1 Question-marks germinating out of basement Modern physics axiomatic is constructed on a very abstract mathematical formalism that is aimed only at the quantitative description of physical phenomena. For instance, the Maxwell equations describe electromagnetic phenomena but they do not bring an idea of the structure of the charge and the electromagnetic field. The Schrödinger equation describes quantum mechanics of particles but the equation cannot explain the reason of long-range action and wave behavior of the particles. On the question what is photon?, quantum electrodynamics answers: it is something that can be described by the equation $$^2\stackrel{}{A}/t^2c^2^2\stackrel{}{A}/\stackrel{}{r}^2=0$$ where $`\stackrel{}{A}`$ is a physical value called the ”vector potential”. Such a structure of the formalism does not permit to reveal the origin of processes constituted the essence of quantum phenomena studied. Lorentz also pointed the same in the beginning of 20th century (see, e.g. Ref. ). De Broglie held the viewpoint that there are hidden laws (see also Bohm ), which provides the basis for motion and that the description of phenomena should also be the goal of physics, not only their prediction. Nonetheless, it is now believed that so-called ”unorthodox” questions are irrelevant. However, can we correctly understand the behavior of elementary particles, if the whole series of fundamental notions researchers operate with everyday have not become clear yet? For example, one can raise the following questions. 1) What is $`\psi `$-wave function? This problem still thrills the curiosity of researchers (see, e.g. review ). 2) All correct theories should be Lorentz invariant, i.e. they and Einstein’s special relativity should agree (see, e.g. Ref. 5). Nevertheless, the Schrödinger equation is not Lorentz invariant but it perfectly describes quantum phenomena. How is it possible? 3) Why does the classical parameter $`M`$ – the particle mass – enter the Schrödinger quantum equation? Where is the particle mass when the particle as the whole is fuzziness in an undetermined volume as the $`\psi `$-wave function prescribes? 4) What is mass? In modern quantum field theories mass is considered as a characteristic expressed through the energy $`E`$ and momentum $`p`$. Today theorists try to assure \[6-8\] that mass does not depend on velocity. And this is very strange because it turns out from such determination that the notion of mass can not be considered as a quantity of matter which is found in the volume of a particle/body. Moreover such declarations are in contradict to the experiments by Bucherer and Rogers et al. who studied the dependence of mass on the velocity and confirmed the validity of the formula $`M=M_0(1v^2/c^2)^{1/2}`$. 5) What are microscopic processes changing the geometry of space surrounding an object, which manifest themselves in the form of the Newton/Coulomb potential $`1/r`$? Einstein noted that the geometry employing in general relativity (the Riemannian geometry) should be treated only as a macroscopic geometry. In other words, what is origin of the gravity? 6) There is no correct determination of values $`E`$ and $`\nu `$ in the expression $`E=h\nu `$ applied to a moving canonical particle. In one case $`E=\frac{1}{2}M_0v^2`$ (see, e.g. Ref. ), and in the other one $`E=M_0c^2(1v^2/c^2)^{1/2}`$ (see, e.g. Ref. ). Which is true? 7) The description of a quantum system in terms of the Dirac field or Dirac’s equation is correct only at the scale $`r\lambda _{\mathrm{Com}}`$ where $`\lambda _{\mathrm{Com}}`$ is the Compton wavelength. Hence, what a physical characteristic of space in the vicinity of particle does the parameter $`\lambda _{\mathrm{Com}}`$ describe? And what approach can make used at the scale $`r<\lambda _{\mathrm{Com}}`$? 8) What is spin? It is one more mystery of the microworld. Quantum field theories define it as an ”inseparable and invariable property of a particle” . That is all. 9) What is nature of the phase transition that turns us from the description of a quantum system based on the Schrödinger equation to that based on the Dirac one? 10) What is nature of the fundamental physical constants $`c`$, $`h`$, and $`G`$? If the value of $`c`$ is constant then why does the experiment register the superluminal velocity (from $`c`$ to $`4.7c`$) ? 11) What is electric charge and why is it fractional in quarks? It is said that the charge is something that is written in quotation marks ; constants of gauge interactions are called the ”charge” as well. Thereby, according to the definition the electric charge is a value that is measured by the elementary electric charge unit $`e`$! 12) What is structure of real space? It is easily seen that answers to these questions cannot be found solely in the framework of power mathematical methods of contemporary physics. Notwithstanding this, the questions are pure physical and we should look for the solvability of all these very urgent problems of fundamental science. We should answer what the canonical particle is? What is its size? And what all the properties mentioned above do mean? ## 2 Confusions of quantum theory Present views on the canonical particle are restricted by the following primitive notion that is not reconcilable with a very difficult and formal mathematical construction, which is applied to the description of the quantum systems behavior. First, experimental data correspond to the length $`l10^{17}`$ cm . At this scale we should imagine a ”black box” that unified our perception about the particle. The box is pasted over various labels, which contain legends like these: the mass is equal to $`M`$, the spin is equal to 1/2, the energy $`E`$ is equal to $`h\nu `$, the charge is equal to $`e`$, and so on. In the case of quarks, things get worse: an isolated quark does not exist and that is why we can talk about the box only resorting to indirect information on the spin, color and so on, that is the notions which need submicroscopic investigations themselves. Second, we should rather substitute the notion of fundamental particle (i.e. our abstract black box) by the more abstract notion of fundamental symmetry . Third, at the atom size the fundamental symmetry is suddenly transformed into the $`\psi `$-wave function or spinor $`\widehat{\psi }`$, as the case requires. The case is a function of the ratio $`v/c`$ where $`v`$ and $`c`$ are the particle and light velocity respectively. It turns out that the $`\psi `$-wave function and spinor $`\widehat{\psi }`$ can be not only considered as the fundamental symmetry parameters but parameters of the particle as well. At the same time special relativity says that the value $`v_0`$ is not absolute and depends on a frame of reference. One may choose such a frame of reference that $`v_0`$ will be very close to $`c`$. This means that the Schrödinger formalism may be easily replaced for the Dirac one and on the contrary, we can choose such frame of reference for a quantum system described by the Dirac equation that the Dirac formalism will smoothly pass to Schrödinger’s. However everybody knows that this is absurdity and moreover at this point an internal inconsistency of the theory comes to light: in the Schrödinger quantum equation the distance between two instantly interacting electrons, as was noted by Ehrenfest , ”can be equal to any quantity of kilometers”. Besides the Schrödinger equation is not Lorentz invariant and therefore formalism based on this equation can not be conjugated with that resting on the Dirac one, whereas both the formalisms are confirmed by experiment perfectly. So such a strange theory we have. On the other hand, there are exact postulates, which directly follow experimental facts. First, there are corpuscles whose behavior similar to wave. Second, the velocity and mass are characteristics of objects; the size of objects contracts in the direction of motion and objects’ masses increase with velocities. Third, two basic quantum mechanical relations are applied for any particle: $`E=h\nu `$ and $`\lambda =h/Mv`$. Forth, each particle has its own limiting length, the Compton wavelength $`\lambda _{\mathrm{Com}}`$, behind of which quantum fluctuations of the vacuum are absent. Fifth, there is a quantum characteristic of the particle called the spin that can contribute the orbital momentum of the particle. Sixth, when the velocity $`v`$ of the particle approaches to $`c`$, the phase transition takes place in the quantum system studied and one should pass from the Schrödinger formalism to the Dirac one. Loud disagreements in quantum theory point towards the need for its improvement. All modification must keep pace with reliable established experimental facts. To solve the problem we should try to study the three following subjects together, which have never been previously considered as the whole: Foundations of quantum mechanics, Foundations of quantum gravity, and Foundations of quantum electricity. But the first point of the study is the structure of the geometry of space and a correct definition of real space. ## 3 Search for a submicroscopic approach There are different approaches to the problem. Among them one can name a new approach by Hofer who has proposed to consider electrons and photons as extended particles which comply with a wave equation; the newest concept offered by Kirilyuk , who has constructed a theory of two fundamental fields, which lead to the universal concept of dynamic complexity and the permanently developing hierarchical structure of the universe. Kiriliuk’s model demonstrates a possibility of the double solution with chaos, which takes into account the deterministic concept of quantum mechanics developed by de Broglie. Based on causal interpretation of quantum mechanics pioneered by de Broglie and Bohm , Roy and Singh have suggested a deterministic mechanics in which the quantum probability densities are simultaneously reproduced as marginal of one positive defined phase space density, which is constant along the trajectory. We shall point out also several other new views on the nature of a vacuum and real space which have appeared \[26-32\] during the last decade. These see a vacuum as a substance and determine matter as deformations of space, Bounias and Bounaly ; in particular, papers studied premises for the existence of an initial cell of space in terms of the topology and the set theory. About some kind of a primordial cell and existence principle was also pointed in Ref. . In high energy physics wave properties of canonical particles are neglected and the behavior of a quantum system is often described drawing an analogy with the lattice model, string model, bubble model, bag model, etc. Thus, in many cases quantum field theory can not preclude ideas and concepts used in condensed media physics. Because of this account must be taken of the microstructure of the vacuum, i.e., real space, at the scale of the order of $`10^{28}`$ cm (at this size electromagnetic, weak and strong interactions come together). It is likely that space at this scale can be simulated as an order/disorder lattice, similar to a solid/liquid, or as a cellular structure, similar to the pack of soft spheres. (It is interesting to note that such view is conceptually close to the re-introduction of some kind of an aether, but the quantum one.) In the author’s works \[34-36\] real space has been simulated in the form of a quantum substance as well and an elementary cell of space – a superparticle – has been offered. The origin of matter, a local space curvature, or deformation, is created when the volume of an initial cell changes. One can consider the local deformation as a corpuscle. Let us look now if the model can explain quantum mechanics phenomena. How can the local deformation that can be treated as a corpuscle whose size is limited by $`10^{28}`$ cm moves similar to a wave and manifests the wave behavior at the atom scale? Of course, it is not an easy problem but it has a solution. The motion of the physical ”point” (corpuscle cell) in entirely packed discrete space must be accompanied by the interaction with the ”points” of space (superparticle cells) giving rise to excitations in neighboring cells. Note that the similar phenomenon occurs in a solid: a particle moving in the solid brings about excitations such as excitons, solitons, etc. As the excitations are associated with the motion of the corpuscle they were called ”inertons”, i.e. just the corpuscle inert mass is responsible for the creation of such kind of space excitations. A portrait of the moving corpuscle is depicted in Fig. 1. When the corpuscle moves it also pulls its deformation coat, in other words, the space crystallite. The crystallite migrates by a relay mechanism: in any place of corpuscle location surrounding cells are made ready as the Figure demonstrates. The crystallite is similar to a shell that screens the corpuscle from degenerate space. Cells have mass inside the crystallite and have not it out. The dynamics of the corpuscle and its inerton cloud was studied in Refs. \[34-36\]. It was shown that the cloud of inertons oscillates around the corpuscle with amplitude $`\mathrm{\Lambda }`$ that can be found using the relationship $$\mathrm{\Lambda }=\lambda c/v_0$$ (1) where $`\lambda `$ is the amplitude of spatial oscillations of the corpuscle, $`c`$ is the speed of light and $`v_0`$ is the corpuscle’s initial velocity. It is quite reasonable to center the notion of corpuscle, i.e., local deformation of space, on the notion of canonical particle. This enables to make a deeper microscopic interpretation of the de Broglie wavelength $`\lambda `$ as amplitude of spatial oscillations of the particle along its path: on the first half of spatial period of oscillations $`\lambda /2`$ the particle emits inertons and its velocity decreases from $`v_0`$ to 0. The emitted inertons gradually retarded by elastic cellular space. Then space returns the inertons to the particle and hence on the second half of the period $`\lambda /2`$ the particle absorbs the inertons and its velocity increases from 0 to $`v_0`$, and so on. It is significant that the inertons emitted ahead of the particle are absorbed behind of it (owing to the difference between the inerton and the particle velocities, $`c>v_0`$). The submicroscopic mechanics of the particle permits to deduce correctly the basic quantum mechanical relations: $$E=h\nu ,\lambda =h/p;$$ (2) here, in our case $`p=Mv_0`$ and $`M=M_0(1v_0^2/c^2)^{1/2}`$. Relations (2) were not obtained previously from any theory; they were only postulated by de Broglie who showed that the Schrödinger equation is their consequence. However, unlike the traditional presentation, the Schrödinger equation gained in paper is Lorentz invariant because it includes time $`t`$ as a natural parameter that is Lorentz invariant by definition. Space covered by the inerton cloud determines the range of the wave $`\psi `$-function action. The cloud of inertons accompanying the moving particle contacts any obstacles around the particle in a distance $`\mathrm{\Lambda }`$ and transmits a respective information to the particle and this is the easiest explanation of the particle diffraction phenomenon. The size of the deformation coat, or the crystallite is equal to the Compton wavelength $`\lambda _{\mathrm{Com}}`$; it is just the parameter that characterizes the relativistic behavior of the particle, in particular, the photon scattering by it. What is inerton? It is a quasi-particle that carries an elementary deformation from cell to cell by a relay mechanism like the Frenkel (molecular) exciton transferring the energy in a molecular crystal. Owing to the comparison of the deformation with mass (see Refs. ) it is reasonable to assume that inertons should substitute for hypothetical gravitons – carriers of the gravitational interaction of general relativity. Indeed, i) inertons have mass (gravitons have not); ii) inertons are a part of any quantum and classical physical system (gravitons were deduced only from the pure classical behavior of objects and these particles can not be introduced in quantum mechanics in principle); iii) inertons can be easily revealed in any physical laboratory by means of many different tools (the existence of gravitons has never been confirmed). The general theory of relativity did not take into account the existence of matter waves, which quantize space at the microscopic scale. Therefore the macroscopic requantization of space that general relativity predicts is highly conjectural. Moreover the relativity lumped together real space with time which is a non-geometric parameter. This is the amalgamation that builds up enormous obstacles on the way to a microscopic consideration of the gravitation phenomena. The problems of the construction of a mathematical space and time emergence in it has recently been raised by Bonaly and Bounias ; they have shown that time is associated with the mapping of intersections of topological spaces. A detailed theory of space conforming to the experimental results in the fields of microscopic and macroscopic phenomena is stated in Ref. . The impact of inertons on the structure of test specimens has been demonstrated in paper (other manifestations of inertons are described in Refs. ). At the same time many other physicists have observed unusual effects, which may be caused by inertons as well. In particular, Europhys. News has reported about one of them : a large group of researchers could observe the electron wave $`\psi `$-function on metal surface. Nonetheless, everybody knows that the wave $`\psi `$-function is just the mathematical function like to the Boltzmann function $`f(\stackrel{}{r},\stackrel{}{\dot{r}})`$, the Hamilton-Jacobi function $`S(\stackrel{}{r},E)`$, etc. All these do not act in real space and only set connections between particle’s parameters. This is why spherical and elliptical images showed in the figures in Ref. should be interpreted as images of inerton clouds surrounding electrons. The concept of cellular degenerate space and the submicroscopic mechanics, which are progressing allow us to disclose many significant details of the microworld and it is valid to say that the first significant result obtained in the framework of the concept is the solution of two difficult problems of nonrelativistic quantum mechanics. First, the theory developed in Refs. removed a very unpleasant conflict that took place between nonrelativistic quantum mechanics and special relativity: Unlike the traditional presentation, the Schrödinger equation gained in paper is Lorentz invariant owing to the invariant time entered in the equation. Second, due to inertons introduced in the quantum system nonrelativistic quantum mechanics no longer suffers from long-range action. Such submicroscopic approach is able to give us great insight into both the submicroscopic structure of canonical particles and the particles dynamics, which are hidden from observation inside the ”black box” that presents an impenetrable barrier to the quantum field theory, string theory, supersymmetry, supergravity, and others. The theory based on the concept of fine-grained degenerate space is a newest one, however, it would be the shortest way to the unified theory of matter. The new concept needs new mathematical ideas, new approaches and a new research methodology. Acknowledgement I am very thankful to Professor Michel Bounias for fruitful discussions and valuable remarks.
no-problem/9908/hep-th9908042.html
ar5iv
text
# References KANAZAWA-99-11 KUCP-0139 Wilson Renormalization Group Equations for the Critical Dynamics of Chiral Symmetry Ken-Ichi Aoki , Keiichi Morikawa , Jun-Ichi Sumi , Haruhiko Terao and Masashi Tomoyose Institute for Theoretical Physics, Kanazawa University Kanazawa 920–1192, Japan Department of Fundamental Sciences, Faculty of Ingegrated Human Studies, Kyoto University, Kyoto 606-8501, Japan Abstract The critical dynamics of the chiral symmetry breaking induced by gauge interaction is examined in the Wilson renormalization group framework in comparison with the Schwinger-Dyson approach. We derive the beta functions for the four-fermi couplings in the sharp cutoff renormalzation group scheme, from which the critical couplings and the anomalous dimensions of the fermion composite operators near criticality are immediately obtained. It is also shown that the beta functions lead to the same critical behavior found by solving the so-called ladder Schwinger-Dyson equation, if we restrict the radiative corrections to a certain limited type. 1. Introduction The chiral symmetry breaking phenomena has been one of the key issues to be understood in the non-perturbative dynamics of gauge theories. The analytical study of this problem has been initiated by the Nambu-Jona-Lasinio (NJL) model , which was introduced as the effective theory with four-fermi interactions. For gauge theories particularly the Schwinger-Dyson (SD) equaitons in the ladder approximation with Landau gauge have been intensitively studied and applied not only to QCD but also to the various models of dynamical electroweak symmetry breaking . In QCD, the ladder SD equation with the running gauge coupling constant, the improved ladder , was found to give good results even qantitatively . However the ladder SD equations are known to suffer from some serious problems, specially the strong gauge dependence and the difficulty to proceed beyond the ladder approximation . On the other hand the Wilson renormalzation group (RG) has been known to offer the powerful method to analyze critical phenomena and has been applied to the various dynamical problems mainly in the statistical mechanics. The so-called exact RG equations , which are the concrete formulation of the Wilson RG in the momentum space, has been recently applied to numerical study of non-perturbative dynamics in field theories. The application to the QCD dynamics has been also considered in this framework . The advantageous features of this method, compared with the SD approach, are that the critical behavior is analyzed directly form the RG eqautions, and that it admits the systematic improvement of approximation by the derivative expansion and truncation of the Wilsonian effective action . Interestingly it is rather recent that the RG method has been applied to the fermi liquid theory of superconductvity , which the NJL model was considered in analogy with. Also it should be noted that the fermi liquid theory of high density QCD was studied by the RG analyses . In this paper we examine the chiral critical dynamics in gauge theories by using the exact RG equations, specially putting our attention to the comparison with the SD approach. There have been known several formulations of the exact RG. Here we simply employ the Wegner-Houghton RG equations , which are derived with sharp momentum cutoff, in the so-called local potential approximation for our present purpose. The analyses with the exact RG equations with smooth cutoff may be performed as well . It will be found that the critical behavior is determined from the beta-functions of the effective four-fermi couplings induced by gauge interaction with remakably simple calculation. The phase boundary and also the anomalous dimensions of the composite operators of fermions near the criticality will be evaluated. Our approximation scheme adopted here is even better than the ladder approximation performed in the SD equations on the critical behavior. Actually, as is seen later, if we make further approximation so as to pick up only a few types of the radiative corrections, then the critical behavior is reduced to be identical to that obtained by solving the ladder SD equation. 2. Scheme of the RG equations The ladder SD equation for the fermion mass function is given in the form of an integral equation, where the momentum integration is carried out with sharp cutoff. In order to see the direct relation between the critical dynamics obtained by the two methods; the SD equation and the RG equation, we consider the exact RG equation defined with sharp cutoff in this paper. There have been known the several formalisms for the exact RG . Here we shall adopt the so-called Wegner-Houghton RGE derived as follows. If we devide the freedom of the quantum field $`\varphi (p)`$ into the higher freqency modes with $`|p|>\mathrm{\Lambda }`$ and the lower frequency modes with $`|p|<\mathrm{\Lambda }`$ by introducing the cutoff scale $`\mathrm{\Lambda }`$ in the Euclidean formalism, then the Wilsonian effective action at this scale, $`S_{\text{eff}}[\varphi ;\mathrm{\Lambda }]`$, may be defined by integrating out the higher frequency modes in the partition function. Namely $$Z=\underset{|p|<\mathrm{\Lambda }_0}{}d\varphi (p)e^{S_0[\varphi ;\mathrm{\Lambda }_0]}=\underset{|p|<\mathrm{\Lambda }}{}d\varphi (p)e^{S_{\text{eff}}[\varphi ;\mathrm{\Lambda }]},$$ (1) where $`S_0`$ denotes the bare action with the bare cutoff $`\mathrm{\Lambda }_0`$. This effective action contains the general operators invariant under the original symmetries in the bare action, for example the chiral symmetry of our present concern. The Wegner-Houghton RGE determines the variation of the Wilsonian effctive action under the infinitesimal change of the cutoff $`\mathrm{\Lambda }`$. For example, the RGE for the $`D`$-dimensional scalar field theory is found to be given exactly as $`{\displaystyle \frac{S_{\text{eff}}}{t}}=DS_{\text{eff}}{\displaystyle \frac{d^Dp}{(2\pi )^D}\varphi _p\left(\frac{2D\eta }{2}p^\mu \frac{^{}}{p^\mu }\right)\frac{\delta S_{\text{eff}}}{\delta \varphi _p}}`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{d^Dp}{(2\pi )^D}\delta (|p|1)\left\{\frac{\delta S_{\text{eff}}}{\delta \varphi _p}\left(\frac{\delta ^2S_{\text{eff}}}{\delta \varphi _p\delta \varphi _p}\right)^1\frac{\delta S_{\text{eff}}}{\delta \varphi _p}\text{tr}\mathrm{ln}\left(\frac{\delta ^2S_{\text{eff}}}{\delta \varphi _p\delta \varphi _p}\right)\right\}},`$ (2) where $`t=\mathrm{ln}(\mathrm{\Lambda }_0/\mathrm{\Lambda })`$ is introduced as the scale parameter. The 1st line of the RGE represents nothing but the canonical scaling of the effective action. While the 2nd line comes from the radiative corrections which correspond to the tree and the 1-loop Feynman diagrams including only the propagators with the momentum of the scale $`\mathrm{\Lambda }`$. In the practical analysis it is inevitable to simplify this RGE by some approximation. Here we shall examine the RGE in the so-called local potential approximation . In this approximation the radiative corrections to any operators containing derivatives are ignored in the RGE (2). Therefore solely the potential part of $`S_{\text{eff}}`$, $`V_{\text{eff}}`$, may be evolved with the shift of $`\mathrm{\Lambda }`$. It should be noted that the wave function renormalization is ignored in this scheme. The RGE for the scalar theory is given explicitly by $$\frac{V_{\text{eff}}}{t}=DV_{\text{eff}}\frac{D2}{2}\varphi \frac{V_{\text{eff}}}{\varphi }+\frac{A_D}{2}\mathrm{ln}\left(1+\frac{^2V_{\text{eff}}}{\varphi ^2}\right),$$ (3) where $`A_D=2/(4\pi )^{D/2}\mathrm{\Gamma }(D/2)`$ is the factor from the momentum integration. This equation offers us a set of the infinitely many beta functions for the general couplings appearing in $`V_{\text{eff}}`$. However it should be noted here that each beta function may be evaluated through just one loop corrections with the general effective interactions. The non-perturbative nature of the RGE is supposed to be maintained by solving the infinitely many coupled renormalization equations. Actually it has been known that this approximated RGE is quite effective in the case of the scalar theories . The generalization of this RGE to include fermions has been also studied in the relation with the triviality-stability bound for the Higgs boson mass in the standard model . Now let us consider to apply this formulation to the massless fermions coupled by gauge interaction. For example we may take the action of the massless QED as the bare action $$S_0=d^4x\left\{\overline{\psi }/\psi +e\overline{\psi }A/\psi +\frac{1}{4}F_{\mu \nu }^2+\frac{1}{2\alpha }(_\mu A_\mu )^2\right\},$$ (4) where $`\alpha `$ is the gauge parameter. As is well known, the gauge invariance is not maintained anymore, once the momentum cutoff is performed. Therefore the generic gauge non-invariant operators are generated in the Wilsonian effective action. Then we must encounter the rather complicated problem to pick up the special RG flows corresponding to the gauge invariant theories in the infinite dimensional coupling space. Recently it has been discussed how to deal with the gauge theories in the framework of the Wilson RG by using the modified Slavnov-Taylor identities . Here, however, we shall simply ignore the corrections to the operators including the gauge fields as well as imposing the local potential approximation as the first step towards the analysis of the chiral symmetry breaking phenomena. Then we may avoid the intriguing problem of the gauge invariance, since any gauge non-invariant operators do not appear in the effective action. This approximation is indeed so rough as to make the beta function of the gauge coupling vanish identically, therefore it cannot be supposed to give any picture of the real dynamics of the gauge theories. However, on the other hand, the so-called ladder approximaton used in the SD approach also totally ignores the vertex corrections as well as the corrections to the gauge kinetic functions. Therefore it would be meaningful to examine the RGE in this scheme in comparison with the SDE’s in the ladder approximation. The effect of the running gauge coupling will be discussed in section 4. Nevertheless this approximation scheme is thus rather crude, it will be seen that the chiral critical behavior may well be described. Actually it will be found that this approximation is even better than the ladder approximation applied for the SD equations. The Wilsonian effctive action to be solved by the RGE in this scheme is now reduced to the form of $$S_{\text{eff}}[\psi ,\overline{\psi };\mathrm{\Lambda }]=d^4x\left\{\overline{\psi }/\psi +V_{\text{eff}}(\psi ,\overline{\psi };\mathrm{\Lambda })+e\overline{\psi }A/\psi +\frac{1}{4}F_{\mu \nu }^2+\frac{1}{2\alpha }(_\mu A_\mu )^2\right\},$$ (5) where $`V_{\text{eff}}(\psi ,\overline{\psi })`$ denotes the general potential composed of the chiral symmetric multi-fermion operators. These multi-fermion operators, which are induced by exchange of the “photon” with higher momentum. The so-called gauged NJL model is often examined in the SD approach and the phase diagram in the two parameter space of the gauge coupling and the four-fermi coupling has been examined . However, in the RG point of view, this coupling space should be regarded as a subspace of the infinite dimensional coupling space of the Wilsonian effective action. It should be noted also that these multi-fermi operators are irrelevant or non-renormalizable, and therefore, are not considered in the perturbative QED. However they cannot be simply discarded in the strong coupling region. It will be seen in the next section that the four-fermi coupling turns out to be relevant near the criticality and plays a crutial role for the critical dynamics of the chiral symmetry breaking. 3. Critical dynamics of the chiral symmetry In this section we examine explicitly the Wegner-Houghton RGE in the approximation discussed in the previous section. The form of the effective potential $`V_{\text{eff}}`$ written in terms of the fermions is found to be restricted into a polynomial composed of the following parity and chiral invariant operators, which are mutually independent; $`𝒪_1=(\overline{\psi }\psi )^2+(\overline{\psi }i\gamma _5\psi )^2={\displaystyle \frac{1}{2}}\left\{(\overline{\psi }\gamma _\mu \psi )^2(\overline{\psi }\gamma _5\gamma _\mu \psi )^2\right\},`$ $`𝒪_2=(\overline{\psi }\gamma _\mu \psi )^2+(\overline{\psi }\gamma _5\gamma _\mu \psi )^2,`$ (6) $`𝒪_3=\left\{(\overline{\psi }\gamma _\mu \psi )(\overline{\psi }\gamma _5\gamma _\mu \psi )\right\}^2.`$ Therefore the 4-fermi part of the effective potential may be written down as $$V_{\text{eff}}(\psi ,\overline{\psi };t)=\frac{G_S(t)}{2\mathrm{\Lambda }^2}\left\{(\overline{\psi }\psi )^2+(\overline{\psi }i\gamma _5\psi )^2\right\}+\frac{G_V(t)}{2\mathrm{\Lambda }^2}\left\{(\overline{\psi }\gamma _\mu \psi )^2+(\overline{\psi }\gamma _5\gamma _\mu \psi )^2\right\}.$$ (7) Hereafter let us call $`G_S`$ the scalar four-fermi coupling and also $`G_V`$ the vector four-fermi coupling. <sup>1</sup><sup>1</sup>1The sign of the scalar four-fermi coupling introduced here follows the conventional one in the literatures. Now let us evaluate the radiative corrections to these four-fermi operators, since it will be found enough to see only these couplings for the purpose of understanding the critical dynamics. By using the propagator of the gauge field in the Landau gauge $`(\alpha =0)`$, the RG equations for the four-fermi couplings are found to be $`{\displaystyle \frac{d}{dt}}g_S=2g_S+{\displaystyle \frac{3}{2}}g_S^2+4g_Sg_V+g_S\lambda {\displaystyle \frac{1}{6}}\lambda ^2,`$ $`{\displaystyle \frac{d}{dt}}g_V=2g_V+{\displaystyle \frac{1}{4}}g_S^2g_V\lambda {\displaystyle \frac{1}{12}}\lambda ^2,`$ (8) where we introduced the rescaled couplings, $`g_S=G_S/(4\pi ^2)`$, $`g_V=G_V/(4\pi ^2)`$, $`\lambda =3e^2/(4\pi ^2)`$. Here we should note that any multi-fermi couplings more than four do not take part in the radiative corrections for the four-fermi couplings owing to the 1-loop nature of the RGE. Therefore we may obtain the RG flows within the 2 dimensional coupling space (or 3 dimensional, if the gauge coupling is also taken into account) irrespectively to other couplings. In Fig.1 the Feynman diagrams representing the one-loop corrections to the four fermi couplings are shown. Let us call the corrections given by the diagrams in the dashed-line box in Fig.1 the “ladder” type, and the others the “non-ladder” type hereafter. If we approximate the RGE by restricting to the “ladder type” correction, then the beta function for the scalar four-fermi coupling is found to be given by $$\frac{d}{dt}g_S=2g_S+2(g_S+\lambda /4)^2,$$ (9) where it is noted that the RGE for the scalar four-fermi decouples from that for the vector four-fermi coupling. Before examining the full RGE’s (8), let us consider the RG flows in the subspace of $`\lambda =0`$. The beta functions tell the “fixed points” at $`(g_S^{},g_V^{})=(0,0),(1,1/8),(4,2)`$. $`(0,0)`$ is the IR trivial fixed point, and $`(1,1/8)`$ is the UV fixed point on the critical surface. <sup>2</sup><sup>2</sup>2 Strictly speaking $`(g_S^{},g_V^{})=(1,1/8)`$ is not a fixed point, since the beta function for the eight-fermi coupling does not vanish there. However it turns out to be the non-trivial fixed point for the space-time dimension of $`2<d<4`$. We do not consider the point $`(4,2)`$, since it seems to be fake due to this approximation. The RG flows are found as is shown in Fig.2. It is seen that there are two phases divided by a critical surface. The chiral symmetry is supposed to be spontaneously broken in the upper region in Fig.2. Then we may realize that the chiral symmetry breaking is caused essentially by the strong scalar four-fermi interaction, not by the vector four-fermi interaction. It is also easy to evaluate the exponents which are important physical quantities in the critical dynamics. By linearlizing the RG equations around the UV “fixed point”, the dimensions of the relevant four-fermi coupling and the irrelevant four-fermi coupling are found to be $`2`$ and $`5/2`$ respectively. The relevant four-fermi operator is given by the combination of $`𝒪_{\text{rel}}=𝒪_1(1/8)𝒪_2`$. The renormalized trajectory in Fig.2 is given by the straight line passing through the non-trivial fixed point. Indeed we may deduce from the RGE’s (8) $$\frac{d}{dt}(g_S8g_V)=(2+\frac{1}{2}g_S)(g_S8g_V),$$ (10) which means that once the combination of $`(g_S8g_V)`$ is vanishing at a point, then it keeps null along the renormalization flow. Therefore the renormalized trajectory is precisely given by the line of $`g_S=8g_V`$. Namely the effective four-fermi operator in the low energy is just $`𝒪_{\text{rel}}`$ irrespectively to the phases. Next let us examine the RG equations (9) given by the “ladder type” corrections. For each gauge coupling $`\lambda `$ there are the UV fixed point and the IR fixed point, which are found to be $$g_S^{}(\lambda )=\left(1\pm \sqrt{1\lambda }\right)^2/4,$$ (11) where $`+`$ is for the UV fixed point and $``$ is for the IR fixed point. Namely they form a fixed line in the $`(\lambda ,g_S)`$ space as shown in Fig.3. The phase boundary is also shown in Fig.3. The upper region is supposed to be the chiral symmetry broken phase. The critical gauge coupling is given by $$\lambda _{\text{cr}}=1.$$ (12) Indeed this critical surface just coinsides with that obtained by solving the SDE in the ladder approximation . The anomalous dimension of the four-fermi coupling $`g_S`$, $`\gamma _G=2+\text{dim}[g_S]`$, near criticality is immediately deduced from the RGE (9) as $$\gamma _G=4g_S^{}(\lambda )+\lambda =2\left(1+\sqrt{1\lambda }\right),$$ (13) which is also found to coincide with the result by the ladder SDE . Therefore it is seen that our approximation used to derive the RGE’s (8) is certainly better than the ladder approximation. Moreover it is easy to take the all corrections shown by Fig.1 including the “ladder type” ones in our framework. It should be noted that the sum of corrections of the “ladder” diagram and the “crossed ladder” given in the last line of Fig.1 is found to be free from gauge parameter dependence. Thus this extention of approximation beyond “ladder” is significant to obtain the gauge independent results . We would like to stress here that the exact RG equations allow us to examine the critical dynamics by remarkably simple calculation, which is a clear contrast with the SD approach. Now we are in a position to go beyond “ladder” by examining the full RG equations (8). The fixed line $`(g_S^{}(\lambda ),g_V^{}(\lambda ))`$, which is given by the solution of the 3rd order equation in turn, is shown in Fig.3 and also in Fig.4 by projection to the $`(\lambda ,g_S)`$-plane and $`(g_V,g_S)`$-plane respectively. It is seen that the critical gauge coupling is now found to be slightly bigger than the value in the “ladder” approximation, ($`\lambda _{\text{cr}}=1.0409`$). In Fig.4 the critical surface separating the two phases is also shown by the cross sections at various gauge coupling up to the critical one. The critical surface given in the case of “ladder” type should be compared with the cross section between the critical surface and the $`g_V=0`$ plane, which is found as shown in Fig.3. It is seen that the phase boundary obtained by our scheme is shifted towards outside compared with the phase boundary, which has been known so far in the ladder SD approach. The exponents at the fixed line also are similarly obtained. The exponent or the dimension of the relevant operator, which was found to be $`2`$ in the previous analysis for $`\lambda =0`$, reduces as the gauge coupling becomes larger. Then it eventually vanishes at $`\lambda =\lambda _{\text{cr}}`$, which is also seen directly from the eq.(8). In Fig.5 the anomalous dimension of the relevant four-fermi coupling $`\gamma _G`$ is presented in comparison with the “ladder” value given by eq.(14). Before ending this section let us mention the anomalous dimension of the fermion mass operator, which we denote $`\gamma _m`$. In order to evaluate it we may incorporate the mass term in the effective action. Then the beta function for the mass $`m`$ may be derived by one-loop diagrams and is found to be $$\frac{d}{dt}m=m\frac{2m}{1+m^2}(g_S+\lambda /4).$$ (14) Here it should be noted that the contribution from the “non-ladder” type corrections vanishes, therefore the vector four fermi coupling does not appear in this beta function. The anomalous dimension on the fixed line is simply given by $$\gamma _m=2g_S^{}(\lambda )+\frac{\lambda }{2},$$ (15) which is shown in Fig.5 as well. In the “ladder” case it is seen that the anomalous dimensions satisfy the relation, $`\gamma _m=\gamma _G/2`$, which has been known also in the analysis of the ladder SDE . In our analysis, however, $`\gamma _m`$ turns out to be fairly larger than $`\gamma _G/2`$ and also than $`\gamma _m`$ obtained so far in the ladder SD approach. 4. RG flow with the running gauge coupling So far the effect of the renormalization of the gauge coupling has been totally ignored, therefore the obtained phase diagrams do not reflect the realistic ones for gauge theories. In the SD approach, the so-called improved ladder approximation , in which the gauge coupling is simply replaced by the running coupling subject to the perturbative RGE apart from the SD framework, has been often used. However this prescription cannot be regarded as systematic improvement of the approximation. On the other hand, in the Wilson RG framework it is possible to include the correction to the gauge coupling naturally by improvement of the previous approximation. This makes a clear contrast to the improved ladder approximation. If we try to treat the non-perturbative dynamics by strong gauge interactions faithfully in the Wilson RG framework, we must encounter the hard problems such as extraction of the gauge invariant theories, development of simple approximation scheme, incorporation of the topological excitations and so on. However, as a primitive approximation, we may evaluate the Wilson beta function of the gauge coupling by the 1st order correction, namely the perturbative one. Then it is enough to solve the RG equations given by (8) in turn coupled with the perturbative RG equation for the gauge coupling $`\lambda `$. In Fig.6 the RG flows for QED obtained in this manner are shown in the $`(\lambda ,g_S)`$-plane. The critical surface separating the spontaneously broken and the unbroken phases is maintained, while the non-trivial fixed points turn out disappear. Note that the point $`(\lambda ,g_S^{},g_V^{})=(0,1,1/8)`$ is not a fixed point in 4 dimensions. The RG flows for the QCD like asymptotically free gauge theory is shown in Fig.7. These results should be compared with those obtained by solving the SD equations . It is seen that the phase structure is completely swept off. The entire reagion is supposed to be in the broken phase of the chiral symmetry, since the 4-fermi coupling keeps growing in the infrared. The effective theories on the attracting line coming out from the trivial fixed point correspond to the continuum limit of QCD. Namely this line gives the so-called renormalized trajectory of QCD. However other flows of the gauged NJL models, specially starting at the critical point of the NJL model in the UV limit, do not converge on the renormalized trjectory. Therefore it may be supposed that these flows show other renormalized trajectories. If this is the case, the gauged NJL model offers non-perturbatively renormalizable theories different form QCD. Indeed existence of the non-trivial continuum limit other than QCD, or renormalizability of some kinds of the gauged NJL models has been claimed so far . In our frame work of the non-perturbative RG, renormalizability of the gauged NJL model may be shown by examining whether these flows are really the renormalized trajectories or not. Such studies will be reported separately . 5. Discussions In this paper we examined the chiral critical behavior of the gauge theories in the Wilson RG framework. We considered evolution of the effective potential composed of the chirally invariant multi-fermi operators by the exact RG equation with sharp cutoff in the local potential approximation. The RG flow of the four-fermi couplings were found to determine the phase structure. It is straightforwards and remarkably easy to find the critical surface and the anomalous dimensions of the composite operators of fermions in this framework. Moreover the critical dynamics obtained by solving the ladder SD equations are exactly reproduced by restricting the radiative corrections taken in the beta functions to the “ladder” type. While our RG equations contain also the “non-ladder” type corrections, which are neccessary to obtain gauge independent physical results , though we have considered only the case of the Landau gauge in this paper. However we cannot assert with this analysis that the chiral symmetry is indeed spontaneously broken in the region supposed to be the broken phase. In order to clarify it we need to evaluate the order parameters such as the dynamical mass of the fermion, the condensation of the composite operator of fermion, and so on. Evaluation of these order parameters is important especially in QCD, because they are the physical quantities to show the dynamical chiral symmetry breaking. While, for example, the spontaneous generation of the fermion mass itself seems to be even non-trivial in the Wilson RG picture, since the Wilsonian effective action remains chiral symmetric in evolution, that is, there is no room for the mass term to show up. These issues as well as the method to evaluate the order parameters in the Wilson RG framework will be also reported in a separate publication . In our analyses the RG eqaution for the gauge coupling was approximated by the perturbative one. Of course the fully non-perturbative treatment for the RGE is required to see the dynamics of strong gauge interaction in infrared. It would be still an open question whether the Wilson RG approach gives a useful framework in this non-perturbative region. However it may be said that the non-perturbative RG has a good chance to seek for the dynamical chiral symmetry breaking phenomena in gauge theories further by going beyond the level examined so far in the SD approach.
no-problem/9908/quant-ph9908071.html
ar5iv
text
# 1 Introduction ## 1 Introduction Quantum physics is a very successful theory for predicting nature. However, in spite of many attempts, a mathematically and philosophically convincing basis for the interpretation of quantum phenomena has not yet been found. As stated recently by Zeilinger , there are at least two levels of interpreting quantum mechanics: the statistical interpretation in the narrower sense introduced by Born , on which there is almost complete consensus between physicists, and the interpretation of the meaning of the quantum mechanical concepts, where no agreement has been reached and all existing interpretations have been found wanting. The ‘orthodox’ Copenhagen interpretation in terms of state reduction by external measurement loses credit because it becomes meaningless for the universe as a whole. There have been a number of proposals to change the structure of quantum mechanics, e.g., through pilot waves (Bohm ), consistent histories (e.g., Omnès ), event enhancement (Blanchard & Jadczyk ), or gravitational objective reduction (Penrose ). Perhaps the main reason why there is so little progress on the meaning of quantum mechanical concepts is that it is usually seen as thoroughly enmeshed with measurement problems. I believe that this is a mistake, and that more clarity can be obtained by separating the analysis of meaning from that of observability. The discussion thereby becomes more concise and clear. Here I refer to meaning as a conceptual, logically consistent mathematical framework that allows one to speak unambiguously about all the terms used in the theory, in a way intuitively related to corresponding concepts of external reality. This part is often elegant and concise. On the other hand, I refer to observability questions (or measurement problems) as the operational explanation of how to obtain quantitative values for the observables of a system, preferably within the framework of a well-defined mathematical theory. This part is usually messy, and it is the part that I propose to avoid until the other part is satisfactory. (This also saves me from entering a discussion about the meaning of words like ‘operational’ or ‘obtain’.) For example, in classical real analysis, infinitesimals are an ill-defined, meaningless concept, though they can be approximately realized, while random sequences are a well-defined, meaningful concept, though there is no constructive way of finding one, except approximately. A historical case supporting the power of this view of separating the analysis of meaning from that of observability is the uncertainty principle. It was argued by Heisenberg in terms of observability, and later by Robertson in mathematical terms as a simple consequence of the canonical commutation relation $`[p,q]=ih^{}`$. The mathematical argument takes only a few lines and is simple and compelling; the discussion of observability is complex but shows that the results of the mathematics cannot be in conflict with what can in principle be measured by experiment. Further support for the positive effect of the separation of meaning from observability is given by Bell’s inequality (Bell , Clauser et al. ), a purely algebraic statement whose clarity is compelling, and the subsequent verification of its violation through experiments by Aspect . Finally, it seems that measurement problems can be adequately analysed by generalized observables defined as positive operator valued (POV) measures; see, e.g., Davies , Busch et al. . The gradual and approximate state reduction can be explained thermodynamically through dissipative interaction by the coupling to a macroscopic apparatus or heat bath; see, e.g., Zurek , Joos & Zeh , Ghirardi, Rimini & Weber . That obtaining information about a system and state reduction are not equivalent things is a consequence of the possibility of nondemolition measurements (Braginsky et al. , Kwiat et al. ) that produce knowledge about a system but (almost) avoid state reduction. And since the measuring apparatus necessarily involves a huge number of particles, any satisfactory solution of measurement problems should not be a part of the foundational concepts and these but be obtained as their consequence. In an attempt to separate the discussion about the meaning of quantum mechanics from that of experimental consistency, I searched the literature for concise statements of the basic difficulties in terms unrelated to measurement, but given in a clear mathematical context that allows one to analyse the logic of the problems without losing oneself in philosophical speculations. The outcome was somewhat surprising: The best presented mathematical arguments turned out to have subtle flaws. This becomes apparent when one reduces them to arguments residing completely within the universally accepted statistical interpretation in the narrower sense. Probably the weaknesses in the arguments escaped previous attention since the arguments were only used for philosophical discussion and justification of foundations, and never subjected to experiment. I therefore suggest below some specific things that one should try to test with suitable experiments. In the following, I shall present my analysis of some arguments related to a realistic interpretation of quantum mechanics; realistic in the sense that it allows to speak of definite values for all observables at any time that are, however, only partially measurable. The findings give rise to the hope that a consistent interpretation is possible that respects the stochastic nature of quantum mechanics, but is also realistic, with all the accompanying advantages for our intuition. In a companion paper (Neumaier ), a proposal for such a realistic interpretation will be presented. ## 2 Feynman’s argument In his 1948 paper “Space-time approach to non-relativistic quantum mechanics” , Richard Feynman gives on pp.368-369 the following argument. “We define $`P_{ab}`$ as the probability that if the measurement $`A`$ gave the result $`a`$, then measurement $`B`$ will give the result $`b`$. \[…\] denote by $`P_{abc}`$ the probability \[…\] if $`A`$ gives $`a`$, then $`B`$ gives $`b`$, and $`C`$ gives $`c`$. \[…\] If the events between $`a`$ and $`b`$ are independent of those between $`b`$ and $`c`$, then $$P_{abc}=P_{ab}P_{bc}.$$ (1) This is true according to quantum mechanics when the statement that $`B`$ is $`b`$ is a complete specification of the state. In any event, we expect the relation $$P_{ac}=\underset{b}{}P_{abc}.$$ (2) \[…\] Now, the essential difference between classical and quantum physics lies in Eq. (2). In classical mechanics it is always true. In quantum mechanics it is often false. \[…\] The classical law, obtained by combining (1) and (2), $$P_{ac}=\underset{b}{}P_{ab}P_{bc}$$ (4) is replaced by $$\varphi _{ac}=\underset{b}{}\varphi _{ab}\varphi _{bc}.$$ (5) \[for probability amplitudes $`\varphi _{ab}`$ with $`|\varphi _{ab}|^2=P_{ab}`$\] If (5) is correct, ordinarily (4) is incorrect. \[…\] Looking at probability from a frequency point of view (4) simply results from the statement that in each experiment giving $`a`$ and $`c`$, $`B`$ had some value. \[…\] Noting that gzite5 replaces (4) only under the circumstance that we make no attempt to measure $`B`$, we are lead to say that the statement, ‘$`B`$ had some value,’ may be meaningless whenever we make no attempt to measure $`B`$.” Feynman’s argument sounds convincing but it has a flaw in the assumption of (1), the conditional independence of $`c`$ from $`a`$, given $`b`$, for all states summed over in (2). This assumption is inconsistent with quantum mechanics. Indeed, one of the characteristic features of quantum mechanics is its inseparability, the impossibility of splitting a quantum mechanical system into two that are independent. According to the Copenhagen interpretation (Bohr ), and verified abundantly by experiment, the whole space-time set-up must be specified to determine the correct outcome of an experiment. The contradiction obtained by Feynman in comparing (4) and (5) therefore need not have the consequence that $`B`$ sometimes has no value, but serves more naturally as a strong argument for quantum inseparability. The Markov property (1) does typically not even hold for consecutive events; due to the inherent noncommutativities, states ”remember” something about their whole history and ”anticipate” something about their whole future. To see this, pick $`a,b,c`$ as observations of the same observable at increasing times (for memory) and at decreasing times (for anticipation). It should be an interesting project to find the conditions under which (and the accuracy to which) the Markov property for a completely observed time series of a sufficiently isolated system, lost in the quantum domain, is recovered in a thermodynamic limit of a large number of particles. We now show in detail that (1) is inconsistent with quantum mechanics. According to the standard interpretation of probabilities as expectations of projection operators we have $`P_{ab}=a|\mathrm{\Pi }_b|a`$ and $`P_{bc}=b|\mathrm{\Pi }_c|b`$, where $`\mathrm{\Pi }_b=|bb|`$ is the projector to state $`b`$ (and similarly for $`c`$). In order that $`P_{abc}`$ makes sense in the standard interpretation we need $`P_{abc}=a|\mathrm{\Pi }_{bc}|a`$ with a projector $`\mathrm{\Pi }_{bc}`$ expressing the proposition $`\mathrm{\Pi }_b\&\mathrm{\Pi }_c`$ that $`B`$ is in state $`b`$ and $`C`$ is in state $`c`$. In orthodox quantum logics (Birkhoff & von Neumann , Svozil ), one considers a projector as representing the subspace of eigenstates to the eigenvalue 1 of the projector and takes the logical operation $`\&`$ as subspace intersection. One finds that, whenever $`\mathrm{\Pi }_b\mathrm{\Pi }_c`$, we have $`\mathrm{\Pi }_b\&\mathrm{\Pi }_c=0`$ and thus $`P_{abc}=0`$. This is different from (1), proving that (1) has nothing to do with quantum mechanics. And when $`\mathrm{\Pi }_b=\mathrm{\Pi }_c`$, we have $`\mathrm{\Pi }_b\&\mathrm{\Pi }_c=\mathrm{\Pi }_c`$, hence $`P_{abc}=a|\mathrm{\Pi }_c|a=P_{ac}`$, and (2) turns out to be always true in the quantum logic calculus, in contrast to what Feynman claimed. On the other hand, the statement $`P_{abc}=0`$ if $`\mathrm{\Pi }_b\mathrm{\Pi }_c`$ seems somewhat counterintuitive, since if $`\mathrm{\Pi }_b\mathrm{\Pi }_c`$ one might want to have $`P_{abc}P_{ac}`$. In view of the fact that, unless $`\mathrm{\Pi }_b`$ and $`\mathrm{\Pi }_c`$ commute, there is no simple algebraic expression relating $`\mathrm{\Pi }_{bc}`$ to $`\mathrm{\Pi }_b`$ and $`\mathrm{\Pi }_c`$, one is not compelled to adhere to the rigid assignment $`\mathrm{\Pi }_b\&\mathrm{\Pi }_c=0`$ and might prefer to consider $`\mathrm{\Pi }_b\&\mathrm{\Pi }_c`$ to be undefined if $`\mathrm{\Pi }_b`$ and $`\mathrm{\Pi }_c`$ do not commute, interpreting this as an indication that $`P_{abc}`$ cannot be predicted (in principle) by the formalism of quantum theory (and thus may have any value). In view of the fact that the counterintuitive abundance of zero projectors has further undesirable consequences to be discussed in Section 5, this would seem a more satisfactory solution of Feynman’s argument. A new quantum logic consistent with this interpretation is discussed in Neumaier . ## 3 Wigner’s claim Similar discussions as those given by Feynman usually suffer from the same problem that the claimed probability cannot be written as the expectation of a quantum proposition, i.e., a projector. For example, in his 1976 lecture notes on the ‘interpretation of quantum mechanics’, Wigner \[32, p.288\] writes, \[…\] it remains essentially correct to say that the basic statement of quantum mechanics can be given in a formula as simple as (42). In adapted notation, for a finite sequence $`a,b,c,d\mathrm{}`$ of results of measurements of $`A,B,C,D,\mathrm{}`$, it reads $$\begin{array}{cc}P_{abcd\mathrm{}}\hfill & =tr\mathrm{\Pi }_a\mathrm{\Pi }_b\mathrm{\Pi }_c\mathrm{\Pi }_d\mathrm{}\mathrm{\Pi }_d\mathrm{\Pi }_c\mathrm{\Pi }_b/tr\mathrm{\Pi }_a\hfill \\ & =a|\mathrm{\Pi }_b\mathrm{\Pi }_c\mathrm{\Pi }_d\mathrm{}\mathrm{\Pi }_d\mathrm{\Pi }_c\mathrm{\Pi }_b|a\text{if }a\text{ is a pure state.}\hfill \end{array}$$ (42) Now there is a problem with (42). It implies that repeating the measurement of $`A`$ immediately after a first measurement of $`A`$ gives the result of the first measurement with certainty. This is possible only if we assume that the measurements performed are ideal. But (cf. Wigner’s discussion in pp.283-284 of ; see also ) one cannot make an ideal measurement of an observable with a continuous spectrum; and the quantum mechanical analysis of the measurement process shows that an ideal measurement of any quantity takes, strictly speaking, an infinite time. And even allowing for that, the only measurable observables would be the functions of scattering invariants (p.298). Hence formula (42) can apply to real situations only approximately. Therefore, unlike Schroedinger’s equation, (42) cannot be considered a ‘basic statement’ and must be banned from the foundations. It should rather be a consequence of more elementary, exact features of the theory. Indeed, Wigner ‘derives’ (42) from first principles. The argument, given on pp.286-287, is identical to Feynman’s, assuming (without comment) the Markov property that provides the independence that allows to multiply probabilities: If the first observation \[…\] This probability is $`|(a_\kappa ,b_\lambda )|^2`$. The probability that the next measurement \[…\] is then $`|(b_\lambda ,c_\mu )|^2`$ and the probability of both outcomes \[…\] is $`|(a_\kappa ,b_\lambda )(b_\lambda ,c_\mu )|^2`$ \[…\]. Thus the formula, based on invalid reasoning, is suspect. This suspicion is confirmed by realizing that the operator whose expectation is taken is generally not a projector, hence does not correspond to a ‘proposition’ in the traditional quantum logic calculus. But what sense should it make to talk about the probability of a statement that is not even logically well-formed? It seems to me that the only formula for probabilities verified (abundantly) by experiments is the formula (see, e.g., equation (1.8) in Chapter 3 of Davies ) $$P(BE|\rho )=tr\rho B(E)$$ (6) for the probability that, in a mixed state with the density matrix $`\rho `$, the generalized observable (POV measure) $`B`$ has a value from a set $`E`$. For ordinary observables $`B`$ and singleton sets $`E=\{b\}`$, the ‘effect’ $`B(E)`$ reduces to a projector $`\mathrm{\Pi }_b`$. Equation (6) contains a special case of (42) only, $$\begin{array}{cc}P_{ab}\hfill & =tr\mathrm{\Pi }_a\mathrm{\Pi }_b/tr\mathrm{\Pi }_a\hfill \\ & =a|\mathrm{\Pi }_b|a=|a|b|^2\text{if }\mathrm{\Pi }_a=|aa|,a|a=1.\hfill \end{array}$$ (7) It is interesting to note that Wigner’s formula (42) reappears as the basic formula in the theories of consistent histories (e.g., equation (4.3) in Omnès ). However, as mentioned on p.143 of , only the case of two reference times can be proved, and this is equivalent to (6) (with $`\rho `$ replaced by $`E_1\rho E_1`$, the mixed state obtained after the first state reduction). However, we can give the probability (42) an alternative, correct meaning: If $`a`$ is a pure state then (6) implies that $`P_{abcd}`$ is the probability of measuring $`\mathrm{\Pi }_a\mathrm{\Pi }_b\mathrm{\Pi }_cD\mathrm{\Pi }_c\mathrm{\Pi }_b\mathrm{\Pi }_a`$ in that state! In general, therefore, Wigner’s formula gives the probability of measuring in state $`a`$ the last element in a finite sequence $`A`$, $`\mathrm{\Pi }_aB\mathrm{\Pi }_a`$, $`\mathrm{\Pi }_a\mathrm{\Pi }_bC\mathrm{\Pi }_b\mathrm{\Pi }_a`$, $`\mathrm{\Pi }_a\mathrm{\Pi }_b\mathrm{\Pi }_cD\mathrm{\Pi }_c\mathrm{\Pi }_b\mathrm{\Pi }_a`$, $`\mathrm{}`$, assuming that the measurements resulted in reductions of the intermediate states to the states represented by the projectors $`\mathrm{\Pi }_a,\mathrm{\Pi }_b,\mathrm{\Pi }_c,\mathrm{}`$. To see how we can possibly interpret this in terms of the uncontroversial part of quantum mechanics, we consider a quantized relative of Laplace’s demon. Suppose there were a quantum demon with the unusual capacity to ‘see’ every detail of a closed system, without interacting with the system. (The demon doesn’t need to be physically realizable; this is just a fictional argument to make a vivid point.) The demon leaves the whole system (consisting of the measured system together with the measuring device) undisturbed and only interprets our claims on measurements and compares it with the dynamics it sees. The demon notes the time intervals $`t,u,v,\mathrm{}`$ between the successive measurements $`A,B,C,D,\mathrm{}`$. What does it see us measure? In the Heisenberg picture, with unlabelled operators at the time of the first measurement, the state of the complete system remains unchanged, and the measurements are those of $`A,B(t),C(t+u),D(t+u+v),\mathrm{}`$. If we use the Heisenberg dynamics and introduce the commuting operators $`\mathrm{\Phi }_a=\mathrm{exp}(itH/h^{})`$, $`\mathrm{\Phi }_b=\mathrm{exp}(iuH/h^{})`$, $`\mathrm{\Phi }_c=\mathrm{exp}(ivH/h^{})`$, we see that the demon sees us measure $`A,\mathrm{\Phi }_a^{}B\mathrm{\Phi }_a,\mathrm{\Phi }_a^{}\mathrm{\Phi }_b^{}C\mathrm{\Phi }_b\mathrm{\Phi }_a,\mathrm{\Phi }_a^{}\mathrm{\Phi }_b^{}\mathrm{\Phi }_c^{}D\mathrm{\Phi }_c\mathrm{\Phi }_b\mathrm{\Phi }_a,\mathrm{}`$. (This is quite different from what Wigner asserts in his equation (43), where he assumes an additional Heisenberg dynamics of the operators between the state-reducing measurements.) Comparing with the above interpretation of (42), and noting that projectors are Hermitian, we see that the difference between the unproved formula (42) and the demon’s objective description of our sequence of measurements is the replacement of the commuting unitary $`\mathrm{\Phi }`$’s in the complete system by the noncommuting idempotent Hermitian $`\mathrm{\Pi }`$’s in the measured system alone. Thus (42) is valid precisely to the extent that the time-dependent operators (in the Heisenberg picture), as calculated by the Heisenberg equation for the complete experimental setup at the time the next measurement is taken, can be approximated in the assumed (time-independent) state of the investigated subsystem by a two-sided multiplication with these projectors. This shows clearly the aim of a correct measurement theory: It must exhibit broad sufficient conditions for a system that guarantee, for all observables to be measured, the validity of this approximation, and hence a complete reduction of the wave packet. I’d like to challenge the adherents of (42) to devise and perform experiments testing validity, accuracy and limits of (42) when more than two (noncommuting) projectors are involved and the measurements are not ideal. The findings should coincide with the interpretation in the Heisenberg picture just given. (For the realization of arbitrary discrete projectors, note that these can always be brought into the form $`UDU^{}`$ with diagonal projectors $`D`$ and unitary $`U`$; hence Reck et al. applies.) On the basis of such experiments it might also be possible to decide whether it is at all possible to perform experiments for measurements of probabilities that cannot be expressed as the expectation of a projector. ## 4 Schrödinger’s argument In his 1936 paper “Die gegenwärtige Situation in der Quantenmechanik” (where the famous cat paradox appears), Erwin Schrödinger gives the following argument (pp. 156-157 of the English translation): At first thought one might well attempt likewise to refer back the always uncertain statements of Q.M. to an ideal ensemble of states, of which a quite specific one applies in any concrete instance – but one does not know which one. That this won’t work is shown by the one example of angular momentum, as one of many. Imagine \[…\] the point $`M`$ to be situated at various positions relative to $`O`$ and fitted with various momentum vectors, and all these possibilities combined into an ideal ensemble. Then one can indeed choose these positions and vectors that in every case the product of vector length by length of normal $`OF`$ \[where $`F`$ is the point closest to $`O`$ on the line through $`M`$ along the momentum vector\] yields one or the other of the acceptable values – relative to the particular point $`O`$. But for an arbitrary different point $`O^{}`$, of course, unacceptable values occur. \[…\] One could go on indefinitely with more examples. \[…\] Already for the single instant things go wrong. At no moment does there exist an ensemble of classical states of the model that squares with the totality of quantum mechanical statements of this moment. \[…\] we saw that it is not possible smoothly to take over models and to ascribe, to the momentarily unknown or not exactly known variables, nonetheless determinate values, that we simply don’t know. The problem is indeed unsolvable if one insists on the existence of a single vector $`J`$ of values determining all linear combinations $`uJ`$ of the angular momentum. For example, in measuring the inner product $`uJ`$ of the angular momentum $`J`$ with four or more triplewise linearly independent unit vectors $`u`$, one obtains in each case values from the discrete spectrum of $`uJ`$, and these spectra are inconsistent with a precise value of $`J`$. This is explained by Schrödinger in detail on p. 164 in the context of a simpler one-parameter example involving the observables $`p^2+a^2q^2`$ where $`a`$ is a parameter. Commenting on it, Schrödinger writes on p. 165: Should one now think that because we are so ignorant about the relations among the variable-values held ready in one system, that none exists, that far-ranging arbitrary combination can occur? That would mean that such a system of ”one degree of freedom” would need not merely two numbers for adequately describing it, as in classical mechanics, but rather many more, perhaps infinitely many. Schrödinger dismisses this as not viable, but as I’ll show now, there is a serious possibility that precisely this is the case, thus invalidating Schrödinger’s conclusion that it is not possible to ascribe to the momentarily unknown or not exactly known variables determinate values. His conclusion is only valid if one ascribes to each observable vector $`K`$ a determinate value that is to be used in the calculations for all $`uK`$. However, as was observed already by von Neumann \[22, IV.1.E\], one cannot, in general, combine the measuring recipes for two noncommuting observables to one for their sum, so that the sum of two observables is only implicitly characterized through the axioms. This should forbid the naive use of values for the components of $`K`$, say, to calculate the values of $`uK`$. As I shall show now, one may consider each $`uK`$ as a (not necessarily classical) random variable in its own right, and determine their relationship for different $`u`$ not by ordinary algebra but by the statistics derived from the standard quantum mechanical recipes. For simplicity, I’ll take for $`K`$ in place of the angular momentum $`J`$ the Pauli spin vector $`\sigma `$ which shows precisely the same problems as $`J`$. For each unit vector $`uR^3`$, the operator $`u\sigma `$ has the simple eigenvalues $`1`$ and $`1`$; thus these are the possible values of $`u\sigma `$. The projector to an eigenstate of $`u\sigma `$ corresponding to the eigenvalue $`1`$ is $`\mathrm{\Pi }_u:=\frac{1}{2}(1+u\sigma )`$, with $`tr\mathrm{\Pi }_u=1`$, and the corresponding projector for the eigenvalue $`1`$ is simply $`\mathrm{\Pi }_u`$. For any unit vector $`vR^3`$, the probability for $`v\sigma `$ having the value $`1`$ when $`u\sigma `$ has the value $`1`$ is then, according to (7), $$\begin{array}{cc}P_{uv}\hfill & =tr\mathrm{\Pi }_u\mathrm{\Pi }_v=tr(1+u\sigma )(1+v\sigma )/4\hfill \\ & =tr(1+(u+v)\sigma +uv+i(u\times v)\sigma )/4\hfill \\ & =\frac{1}{2}(1+uv),\hfill \end{array}$$ (8) and the probability for $`v\sigma `$ having the value $`1`$ when $`u\sigma `$ has the value $`1`$ is therefore $`P_{u,v}=\frac{1}{2}(1uv)`$, adding up to $`1`$, as it should be. Moreover, $`P_{u,u}=0`$, as it should be. The probabilities depend continuously on $`u`$ and $`v`$, in a very natural way. One would get precisely the same probabilities if one had extremely fragile, classical spheres, painted white on one hemisphere and black on the other one, and each sphere would be destroyed after observing the color at a single point. One could still calculate probabilities for the colors of the other points, and any rotationally symmetric classical probability model produces precisely the same probabilities as we just calculated for the spin. More details on such a classical model for electron spin can be found in Kochen & Specker . They show that, under natural assumptions, such a classical model is restricted to two-state systems. (In a companion paper , I’ll show how one can reinterpret quantum mechanics in such a way that the assumptions of Kochen and Specker become irrelevant, thus removing this very restictive conclusion.) In any case, one such model is enough to invalidate the cogency of Schrödinger’s conclusion. To summarize, Schrödinger’s argument demonstrates that, in quantum mechanics, one cannot calculate the values for linear combinations of noncommuting observables from the values of the observables themselves; but this is already obvious from the properties of the spectrum of operators. Only the values of functions of commuting observables (and, in particular, of functions of a single observable) can be predicted with certainty from the values of the observables themselves. However, as was shown for the case of linear combinations of the Pauli spin matrices, his argument does not demonstrate that one cannot assign natural values to all observables that lead to natural and consistent classical probabilities. The nature of quantum observations may put severe limits on what is observable through experiment and how these are combined to estimate the values of other observables, but it does not seem to put restrictive limits on the joint values of noncommuting observables. Indeed, von Neumann shows in the same section mentioned above how to get estimates of joint probabilities of noncommuting quantities if sufficiently large ensembles are available. ## 5 Particle paths It is widely believed that it is impossible to ascribe definite paths to moving quantum particles, except at special points where the position has been measured. The particle paths seen in a cloud chamber, say, can still be explained as an illusion created by correlation patterns among the atoms ionized by the particles; see Mott . No such analysis seems to be available, however, that would explain why we see macroscopic bound states (such as you or me) move along fairly definite paths. The traditional context for discussing the inconsistency of definite particle paths is the double slit experiment (Bohr ), where it is claimed that electrons (or photons) passing a diaphragm containing the two slits cannot be said to have passed through one or the other slit, or – even worse – are said to have passed through both, in a way (superposition) that defies intuition. Unfortunately, the discussion seems always to be connected to measurement questions, so that it is difficult to discern the mathematical content of the arguments that claim nonexistence of the path (in contrast to its limited measurability only, Wootters & Zurek ). However, I gathered some indirect evidence about particle paths by reading between the lines of some papers. Feynman , in laying ground work for his (nonrelativistic) path integral formalism, discusses on p. 371 the probability of particle paths as his first postulate: If an ideal measurement is performed to determine whether a particle has a path lying in a region of space-time, then the probability that the result will be affirmative is the absolute square of a sum of complex contributions, one from each path in the region. This postulate is used to show (among other things) that the most likely paths are those close to the path determined by the least action principle. No comment is given on the precise mathematical meaning of the terms involved (except for the term “each”); thus we need to see whether we can interpret this in an orthodox way. In the spirit of the standard interpretation of quantum mechanics, we need to assign to each (let us say open and bounded) region $`\mathrm{\Omega }`$ of space-time a projector $`\mathrm{\Pi }_\mathrm{\Omega }`$ to the subspace $`H_\mathrm{\Omega }`$ of all wave functions belonging to states where the particle is with certainty in $`\mathrm{\Omega }`$. Unfortunately, particle positions at different times are represented by operators $`x(t)=\mathrm{exp}(iHt)x(0)\mathrm{exp}(iHt)`$ that generally do not commute with each other, with the consequence that it is unlikely that they have common eigenvectors, which would be the elements of $`H_\mathrm{\Omega }`$. Thus, it is very likely that, for most $`\mathrm{\Omega }`$, this subspace consists of zero only; and $`\mathrm{\Pi }_\mathrm{\Omega }`$ would vanish. This casts serious doubt on the applicability of the quantum logic recipe to defining the projector $`\mathrm{\Pi }_\mathrm{\Omega }`$. Perhaps it is even impossible to make consistent assignments of all projectors $`\mathrm{\Pi }_\mathrm{\Omega }`$, in view of the fact that alternative attempts in consistent histories interpretations run into contradictions (Kent ). (But possibly these contradictions are due to the non-projector nature of these histories; cf. Isham .) However, we can be more modest and only look at special sets of paths for which we may define proper projectors. For example, we may define the distance of two paths $`\xi ^k:[\alpha ,\omega ]R^3`$ ($`k=1,2`$) by $$d(\xi ^1,\xi ^2)=\sqrt{_\alpha ^\omega (\xi ^1(t)\xi ^2(t))^2𝑑t}.$$ (9) Then, for each specific path $`\xi `$, we can use the position operators $`x(t)=\mathrm{exp}(iHt)x(0)\mathrm{exp}(iHt)`$ to define the observable $`\mathrm{\Delta }_\xi :=d(x,\xi )`$ that measures the distance of the particle path from $`\xi `$. Suitable projectors are now the projectors $`\mathrm{\Pi }_{\xi ,\epsilon }`$ projecting to the subspace spanned by the wave functions corresponding to states in which $`\mathrm{\Delta }_\xi `$ has with certainty a value not exceeding $`\epsilon `$. Then, for a system in an arbitrary pure state $`a`$, $$P(d(x,\xi )\epsilon )=a|\mathrm{\Pi }_{\xi ,\epsilon }|a$$ (10) is, according to the most orthodox interpretation of quantum mechanics, a well-defined quantum mechanical probability for the event that the particle path is, in the root mean square sense, within $`\epsilon `$ of the path $`\xi `$. Note that, in principle, this formula can be tested experimentally; and by varying $`\epsilon `$, the complete distribution of $`d(x,\xi )`$ can be found. The natural path to be used for $`\xi `$ is the expectation $$\xi (t)=a|x(t)|a$$ (11) or a computable approximation to it (such as the classical path of the particle). With formulas such as (9)–(11) available, it would be completely unreasonable to assume that a particle has no definite path. If it hadn’t, what should be the meaning of its expectation (11) and of the probability (10) of it being close to a given path!? It remains to discuss how this finding can be reconciled with the double slit experiment. In an implementation of von Neumann’s method mentioned in Section 4 to measure probabilities for two noncommuting observables, Wootters & Zurek measure in a modified double-slit experiment both position and momentum of particles passing a slit: If the measured momentum is positive, then we will guess that the photon passed through slit $`A`$; if it is negative, then we will guess that the photon passed through slit $`B`$. Clearly some of our guesses will be wrong – there are photons that have positive values of measured momentum even though their actual momentum was negative and they went through slit $`B`$. \[…\] these two measurements cannot both be performed for the same photons, and so \[the figures\] cannot refer to the same experiment. Hence, in accord with the Copenhagen interpretation of quantum mechanics, there is no paradox. The complementarity principle does not prevent photons from behaving once as waves and once as particles. It only states that the same photon should not reveal this ”split personality” in the same experiment. \[…\] Despite the fact that we know with $`99\%`$ certainty the paths of the photons, they still have strong wavelike properties. \[…\] we have presented a result which, although not paradoxical, was nevertheless surprising (that is, that one can make a fairly precise determination of the slit through which each photon comes with only a slight disturbance of the interference pattern). The finding loses its surprise in the light of the above considerations that suggest existence of the path but with intrinsic limitations on its observability. For the original double-slit experiment, the arguments generally agree on the observation that there is no way to predict through which slit a particle will pass. However, it is possible to compute with high confidence through which a particle has passed when it has been observed at the photographic plate. To do so, one only needs to compute the probabilities (10) for two typical paths $`\xi `$ connecting each slit with the position where the particle was absorbed by the photographic plate. By comparing the likelihood ratios at a given significance level one will find an exceedingly high likelihood for most particles having passed through the slit closer to their recorded position on the photographic plate. Thus, after the event, enough information is available to decide reasonably reliably on the particle path. One could also calculate the values of the actions of the paths through the two slits ending at the observed position, and use the stationary phase approximation of the Feynman path integral to get (approximations to) the required probabilities. In any case, the statistical analysis is essentially similar to that used for all experiments where an event must be reconstructed from indirect measurable information and from the way the experiment was prepared. ## 6 Conclusion The strongest arguments against realism available in the literature dissolved under a scrutinized analysis, using only that part of quantum mechanics for which there is almost universal agreement about its validity. This opens the door for an interpretation that, while respecting the indeterministic nature of quantum mechanics, allows to speak of definite but only partially measurable values for all observables at any time. This kind of realism is consistent with the intrinsic indeterminism required by Heisenberg’s uncertainty relation if we distinguish carefully between what is and what is measurable. Though not proved by the present investigations, it appears that, independent of the detailed description, a cautious realistic interpretation of quantum mechanics is in full accord with the generally accepted quantum mechanical formalism. Here realistic means that in a completely specified state (not to be confused with the ‘pure states’ of quantum mechanics) all observables have definite values at all times that are, however, only partially measurable, according to the stochastic predictions of quantum mechanics. And cautious means, that one has to take into account the following four restrictions: 1. Only probabilities defined by orthogonal projectors (for sharp observables) or POV measures (for smeared observables) via equation (6) are measurable and predictable. 2. For noncommuting projectors, the logical operation ‘and’ is not defined; the corresponding questions may be asked (and may have definite, i.e., logically consistent answers) but a precise answer cannot be found by experiments. 3. From the values of commuting observables one can deduce values of functions of these observables, but from the values of noncommuting observables one cannot even deduce values of their linear combinations. Instead, functions of noncommuting observables must be considered as random variables on their own whose expectation values (and probability distributions) must be calculated from the general quantum mechanical formalism and not from a classical inference of measured linear combinations. 4. Due to quantum inseparability, the Markov property for a completely observed time series is lost. A constructive proposal for such an interpretation will be given in a companion paper . ## Appendix: Some final speculations In this section I want to be more speculative, at the risk of being less precise, less cogent, and more vulnerable about some issues discussed. I shall mention some ideas and ‘conclusions’ that are related to the preceding analysis though far from being consequences of the rigorous observations of the preceding sections. A realistic interpretation of quantum mechanics is independent of the measurement problem and in better accord with classical (say, 19th century) intuition. As we have seen, there seem to be no longer strong arguments against a cautious realistic interpretation of quantum mechanics. A problem reamining, and one that obscured for a long time the underlying simplicity, is the lack of observable information on the quantum level. As Schrödinger puts it (p.159 of the English translation), referring to the complementarity between position and momentum, \[…\] the momentary statement content of the $`\psi `$-function is far from complete; it comprises only about $`50`$ percent of a complete description. However, unlike Schrödinger, I conclude that, while the remaining 50 percent will probably always remain unobservable, they may have a reality just as objective as the (ideally) observable 50 percent. The fact that in many cases we can choose which 50 percent of a complete description we want to observe underlines this conclusion. The fact that both past and future boundary conditions are needed (and sufficient) to locate a particle path is consistent with the observation in Section 2 that states anticipate something about their whole future. It is also reminiscent of the action-at-a-distance formulation of classical electrodynamics by Wheeler & Feynman , where electromagnetic radiation is explained through interaction with particles in the future. By observing the right $`50\%`$, namely past and future position boundary conditions about one of a pair of conjugate variables at both boundaries, we might be able, in principle, to reconstruct the complete intermediate picture as reliably as in the classical case, where one has the same option. But in the classical case one has the additional and, for us as subjects acting based on past information only, more useful option to predict the particle path from initial conditions only (position and momentum in the past). This state of affairs can be summarized in the statement: Physics essentially describes nature as if everything had already happened, and then expresses its laws as information about observed correlations. Since some of the correlations involve time, it is possible to partially predict the future from the past, or the past from the future, or an intermediate situation from past and future observations. This summary also explains neatly why questions such as the flow of time or free will cannot be discussed within the framework of physics. Whether or not time flows, whether or not our will is free, the four-dimensional picture resulting from the course of nature, whether or not influenced by us, can (in a gedankenexperiment) be replayed, after everything has happened, like a movie. In the replay, everything is determined, and there is only the illusion of free will, just as we are used from the cinema. But the physics, expressed in the correlations between the parts of the movie, is identical to that in the original version. The principle of physics, that it restricts attention to that which remains the same after everything happened, makes physics very powerful in that it allows us to investigate a past of billions of years and to anticipate a future of billions of years. But the same principle also generates its intrinsic limitations, that physics must be silent about everything that cannot be captured by this static, four-dimensional view of nature. This includes both questions such as “how does reality happen?”, and many of the subjects of most interest to people: freedom, purpose, and consciousness.
no-problem/9908/hep-th9908195.html
ar5iv
text
# Gauge Field Theories on a ⟂ lattice ## Introduction The main subject of these notes are difficulties associated with the formulation of gauge field theories on a transverse ($``$) lattice using light-front (LF) quantization. Because of these difficulties, the reader may wonder about the advantages of this approach — particularly given the successes of Euclidean lattice gauge theory (LGT). The primary motivations for formulating QCD in this framework is that LF quantization is the most physical approach towards a microscopic description of the parton distributions measured in deep-inelastic scattering as well as many other hard processes. <sup>1</sup><sup>1</sup>1This and other motivations are discussed in more detail in Ref. mbadv and in references therein. It is important to emphasize this fact in this brief introduction since it explains why LF quantization of QCD and the $``$ lattice should be investigated as a possible alternative to Euclidean and Hamiltonian LGT formulations — despite the difficulties that will be discussed in the remainder of these notes. ### Why LF gauge? Although the choice of quantization hyperplane and the choice of gauge are in principle independent issues, the so-called LF gauge ($`A^+=0`$) turns out to be highly preferable for the canonical formulation of LFQCD. The main reason is that in the kinetic energy term for $`\stackrel{}{A}_{}`$ (from $`\frac{1}{4}F^{\mu \nu }F_{\mu \nu }`$) $$_{kin,A_{}}=D_+\stackrel{}{A}_{}D_{}\stackrel{}{A}_{}=\left(_+igA_+\right)\stackrel{}{A}_{}\left(_{}igA_{}\right)\stackrel{}{A}_{},$$ (1) the term multiplying the ‘time’ derivative of $`\stackrel{}{A}_{}`$ (i.e. $`_+\stackrel{}{A}_{}`$) contains also $`A_{}=A^+`$. Therefore, the canonical momenta $$\mathrm{\Pi }=\frac{}{(_+A_{})}=\left(_{}igA_{}\right)A_{},$$ (2) which are the LF analog to $`\mathrm{\Pi }=\frac{}{(_0A_{})}`$ in equal time quantization, are “simple” (i.e. linear in the fields) if and only if $`A_{}=A^+=0`$. Therefore, in order to avoid having to deal with a system that has to satisfy nonlinear constraints <sup>2</sup><sup>2</sup>2Eq. (2) is a constraint equation since it involves no time-derivative. one normally selects $`A^+=0`$ gauge before quantizing in LF coordinates. However, this choice of gauge is not entirely free of problems. To illustrate this fact, let us start from the Euler-Lagrange equation for $`A^{}`$ in QED <sup>3</sup><sup>3</sup>3Since the problem that we are going to discuss occurs already in QED, we will discuss it there because of the simpler algebra. $$_{}^2A^{}=gJ^++_{}\stackrel{}{}_{}\stackrel{}{A}_{}g\stackrel{~}{J}^+$$ (3) (the LF analog to the Poisson equation), which is also a constraint equation. It is convenient to eliminate $`A^{}`$, using the solution to Eq.(3), i.e. $`A^{}(x^{},\stackrel{}{x}_{})=\frac{g}{2}_{\mathrm{}}^{\mathrm{}}𝑑y^{}\left|x^{}y^{}\right|\stackrel{~}{J}^+(y^{},\stackrel{}{x}_{}),`$ yielding an instantaneous interaction term $$V^{inst}=\frac{g^2}{4}_{\mathrm{}}^{\mathrm{}}𝑑x^{}_{\mathrm{}}^{\mathrm{}}𝑑y^{}d^2x_{}\stackrel{~}{J}^+(x^{},\stackrel{}{x}_{})\left|x^{}y^{}\right|\stackrel{~}{J}^+(y^{},\stackrel{}{x}_{}).$$ (4) This linearly rising interaction in the LF Hamiltonian causes IR divergences, unless $$_{\mathrm{}}^{\mathrm{}}𝑑x^{}\stackrel{~}{J}^+(x^{},\stackrel{}{x}_{})=0\stackrel{}{x}_{}.$$ (5) The origin of this problem lies in the fact that setting $`A^+=0`$ does not completely fix the gauge freedom. An $`x^{}`$ independent gauge transformation $`A^\mu 𝒰^{}A^\mu 𝒰\frac{i}{g}𝒰^{}^\mu 𝒰,\text{with}𝒰=𝒰(\stackrel{}{x}_{}),`$leaves $`A^+=0`$ unchanged and Eq. (5) is just the Gauß’ law constraint associated with this residual gauge symmetry<sup>4</sup><sup>4</sup>4Fixing the remaining gauge freedom requires dealing with explicit zero-mode degrees of freedom and it is still not completely understood how to do this!. As long as $`\stackrel{}{x}_{}`$ is a continuous variable, Eq. (5) implies an infinite number of constraint on the states ($`\mathrm{}`$ number of $`\stackrel{}{x}_{}`$!), which is again difficult to deal with. This is one of the motivations for discretizing the $``$ space direction in the context of LF quantization. ## The transverse lattice The basic idea behind the $``$ lattice bardeen is to work in two continuous ($`x^0`$ and $`x^3`$ or $`x^+`$ and $`x^{}`$) space time directions and two discrete \[$`\stackrel{}{x}_{}(x_1,x_2)`$\] space directions, i.e. space-time consists of a 2-dimensional array of 2-dimensional sheets. <sup>5</sup><sup>5</sup>5In the closely related Hamiltonian LGT space-time consists of a 3-dim. array of 1-dim. lines. The motivation for working with such a ‘hybrid’ formulation is that the discretized $``$ directions provide a the possibility to introduce a gauge invariant cutoff, while the continuous longitudinal directions allow to maintain manifest longitudinal boost invariance (which is one of the advantages of the LF formulation). The natural way to introduce gauge fields within this framework seems to be to work with compact link-fields $`U_{}SU(N_C)`$ in the discretized $``$ directions (as is done in conventional LGT) and with non-compact gauge fields $`A^\pm `$ in the continuous longitudinal directions. It should be emphasized that both the $`U_{}`$’s as well as the $`A^\pm `$ (which are defined on the links and sites of the $``$ lattice respectively), are functions of two discrete and two continuous variables, i.e. one can think of the $``$ lattice action as consisting of many 1+1 dim. gauge theories coupled together. The trouble with this formulation is the nonlinear $`U_{}SU(N_C)`$ constraint on the link fields. The reason that this constraint is more difficult to handle on the $``$ lattice than in Euclidean or Hamiltonian LGT is due to the fact that the $`U_{}`$s are still two dimensional fields (and not just variables, as in Euclidean LGT, or quantum mechanical rotors, as in Hamiltonian LGT). Despite several attempts in this direction paul , nobody has been able to construct a Fock space basis out of these “nonlinear $`\sigma `$ model” degrees of freedom which still allows one to evaluate matrix elements of the LF Hamiltonian. Two possibilities to avoid the problems associated with the $`SU(N)`$-constraint have been pursued: The first is to work with non-compact gauge fields also in the $``$ direction and the other is to keep compact fields, but to relax the $`SU(N)`$ constraint and linearize the degrees of freedom. ### Non-compact formulation of the $``$ lattice Again, we illustrate the main difficulties in the context of QED. In order to satisfy the $`U(1)`$ constraint, one starts with the ansatz $`U_{}=\mathrm{exp}\left(ieA_{}\right)`$, yielding $$P^{}\frac{1}{4}\underset{\stackrel{}{n}_{}}{}𝑑x^{}𝑑y^{}\stackrel{~}{j}^+(x^{},\stackrel{}{n}_{})\left|x^{}y^{}\right|\stackrel{~}{j}^+(y^{},\stackrel{}{n}_{})+P_{plaq}^{},$$ (6) where $`P_{plaq}^{}`$ is the $``$ plaquette interaction ($`xy`$ orientation) and $$\stackrel{~}{j}^+(x^{},\stackrel{}{n}_{})=j_q^+(x^{},\stackrel{}{n}_{})+\frac{1}{e}\mathrm{\Delta }_{}_{}A_{},$$ (7) where $`\mathrm{\Delta }_{}`$ is the discrete approximation to the $``$ Laplace operator and $`j_q^+`$ is the portion of the current carried by the fermions. As long as one restricts oneself to gauge fields with local fluctuations only (as in the Fock expansion!) one finds $`A_{}(+\mathrm{})=A_{}(\mathrm{})`$ and thus $$_{\mathrm{}}^{\mathrm{}}𝑑x^{}\stackrel{~}{j}^+(x^{},\stackrel{}{n}_{})=_{\mathrm{}}^{\mathrm{}}𝑑x^{}j_q^+(x^{},\stackrel{}{n}_{}).$$ (8) Together with Gauß’ law this implies that $`_{\mathrm{}}^{\mathrm{}}𝑑x^{}j_q^+(x^{},\stackrel{}{n}_{})=0`$, i.e. charges must add up to zero at each site! Transversely separated charges are only allowed in the presence of “soliton-like” gauge fields with $`A_{}(x^{}=+\mathrm{})A_{}(x^{}=\mathrm{})=ke`$, where $`k`$ is an integer! The physics behind this result becomes clear by noting that $`\mathrm{exp}\left(ieA_{}\right)`$ acts like such a soliton operator: from the canonical commutation relations $`[A_{}(x^{}),A_{}(y^{})]=\frac{i}{2}\epsilon (x^{}y^{})`$, together with $`[e^A,B]=[A,B]e^A`$ if $`[A,B]`$ is a c-number, one finds $$[A_{}(x^{}=+\mathrm{})A_{}(x^{}=\mathrm{}),\mathrm{exp}\left(ieA_{}(y^{})\right)]=e\mathrm{exp}\left(ieA_{}(y^{})\right).$$ (9) Therefore the Gauß’ law constraint is satisfied if transversely separated charges are separated by a string of exponentials — just as one would have expected from gauge invariance — and the good news is that the infrared divergences cancel if states are constructed in a gauge invariant way. It is instructive to examine in detail how the $`k^+0`$ divergences cancel in $`QED_{2+1}`$. In 2+1 dimensions, there is only one $``$ direction and therefore purely $``$ plaquette terms are absent. As a result, the whole dynamics of $`A_{}`$ is described by its coupling to $`A^{}`$ and pure gauge, coupled to external sources, becomes exactly solvable in the non-compact formulation. The rest frame energy of an external source $`j^+`$ in $`QED_{2+1}`$ on a $``$ lattice is given by $`H_{RF}=v^+P^{}+H_{recoil},`$ where $`v^+`$ is the velocity of the source and $$P^{}=\frac{e^2a}{2}\underset{n}{}_{\mathrm{}}^{\mathrm{}}𝑑q^+\stackrel{~}{j}_n(q^+)\frac{1}{q_{}^{+}{}_{}{}^{2}}\stackrel{~}{j}_n(q^+).$$ (10) is the instantaneous interaction arising from eliminating $`A^{}`$ and $$H_{recoil}=\frac{1}{2v^+}\underset{n}{}_0^{\mathrm{}}𝑑k^+a_n^{}(k^+)a_n(k^+)k^+$$ (11) is a recoil term which appears in the LF description of fixed sources mbzako . The effective current $`\stackrel{~}{j}^+`$ receives contributions from both $`A_{}`$ and the external current $`j^+`$ \[see also Eq.(7)\]. In momentum space, one finds for the current on the $`n^{th}`$ site $$\stackrel{~}{j}_n^+(q^+)=ej_n^+(q^+)+\frac{iq^+}{a}\left[A_{,n}(q^+)A_{,n1}(q^+)\right],$$ (12) where we define the $`n^{th}`$ site to be the one between the $`(n1)^{th}`$ and the $`n^{th}`$ link. It is instructive to decompose the instantaneous interaction into terms quadratic in $`A_{}`$ and $`j^+`$ respectively and a mixed term, i.e. $`P^{}=P_{AA}^{}+P_{jj}^{}+P_{JA}^{}`$, where $`P_{AA}^{}`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk^+}{2a^2k^+}}[a_n^{}(k^+)a_{n1}^{}(k^+)]\left[a_n(k^+)a_{n1}(k^+)\right]`$ (13) $`=`$ $`{\displaystyle \frac{4}{a^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk^+}{2k^+}}{\displaystyle _{\pi /a}^{\pi /a}}𝑑k_{}\mathrm{sin}^2\left({\displaystyle \frac{ak_{}}{2}}\right)a^{}(k^+,k_{})a(k^+,k_{}).`$ where $`a_n(q^+)=_{\pi /a}^{\pi /a}\frac{dq_{}}{\sqrt{2\pi }}a(q^+,q_{})\mathrm{exp}\left(iq_{}an\right)`$. Furthermore $$P_{jj}^{}=\frac{e^2a}{2}\underset{n}{}_{\mathrm{}}^{\mathrm{}}𝑑q^+j_n(q^+)\frac{1}{q_{}^{+}{}_{}{}^{2}}j_n(q^+)$$ (14) is the self-energy of the source, and the coupling of the source to $`A_{}`$ reads $`P_{jA}^{}`$ $`=`$ $`ie{\displaystyle \underset{n}{}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dq^+}{q^+}}j_n(q^+)\left[\stackrel{~}{A}_n(q^+)\stackrel{~}{A}_{n1}(q^+)\right]`$ $`=`$ $`{\displaystyle \frac{ie}{\sqrt{2a}}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq^+}{q_{}^{+}{}_{}{}^{3/2}}}\left\{j_n^+(q^+)\left[a_n(q^+)a_{n1}(q^+)\right]j_n^+(q^+)\left[a_n^{}(q^+)a_{n1}^{}(q^+)\right]\right\}.`$ In order to calculate the self-energy of an external charge-distribution $`j_q(q^+)`$ to order $`e^2`$, one needs to add the instantaneous self-interaction (10) \[which is of $`𝒪(e^2)`$ already\] in first order to the contribution from the coupling to $`A_{}`$ (Non-compact formulation of the $``$ lattice) \[which is only $`𝒪(e)`$\] treated in $`2^{nd}`$ order perturbation theory. The latter yields $`\delta E^{(2)}`$ $`=`$ $`{\displaystyle \frac{v_{}^{+}{}_{}{}^{2}e^2}{a^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dq^+}{q_{}^{+}{}_{}{}^{3}}}{\displaystyle _{\pi /a}^{\pi /a}}𝑑q_{}{\displaystyle \frac{2\mathrm{sin}^2\left(\frac{aq_{}}{2}\right)\stackrel{~}{j}(q)\stackrel{~}{j}(q)}{\frac{q^+}{2v^+}+\frac{2v^+}{a^2q^+}\mathrm{sin}^2\left(\frac{aq_{}}{2}\right)}},`$ (16) where we used the shorthand notation $`j(q)j(q^+,q_{})`$ and where $`j_n(q^+)=_{\pi /a}^{\pi /a}\frac{dq_{}}{\sqrt{2\pi }}j(q^+,q_{})\mathrm{exp}\left(iq_{}an\right)`$. In general, $`\delta E^{(2)}`$ behaves for $`q^+0`$ like $$\delta E_{div}^{(2)}=\frac{v^+e^2}{2}_{\mathrm{}}^{\mathrm{}}\frac{dq^+}{q_{}^{+}{}_{}{}^{2}}_{\pi /a}^{\pi /a}𝑑q_{}j(q)j(q)=\frac{v^+e^2a}{2}_{\mathrm{}}^{\mathrm{}}\frac{dq^+}{q_{}^{+}{}_{}{}^{2}}\underset{n}{}j_n(0)j_n(0),$$ (17) which diverges unless the net charge $`j_n(0)`$ on each (!) site is zero. However, a similar divergence (with opposite sign) arises from the instantaneous interaction, as can be directly read off from Eq. (14). The sum of the two terms is IR finite as long as the total (i.e. sum of charge on all sites) charge is zero<sup>6</sup><sup>6</sup>6This is not surprising since $`QED_{2+1}`$ confines. $`\delta E`$ $``$ $`v^+P_{jj}^{}+\delta E^{(2)}={\displaystyle \frac{e^2a}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dq^+}{v^+}}{\displaystyle _{\pi /a}^{\pi /a}}𝑑q_{}{\displaystyle \frac{\left(\frac{a}{2}\right)^2j(q)j(q)}{\left(\frac{aq^+}{2v^+}\right)^2+\mathrm{sin}^2\left(\frac{aq_{}}{2}\right)}}.`$ (18) In particular, for the case of two (oppositely charged) point charges one finds in the limit $`a0`$ the logarithmic interaction energy, characteristic for an Abelian gauge theory in $`2+1`$ dimensions $$\delta E=\frac{e^2}{2\pi }\mathrm{log}\sqrt{R_{}^2+R_L^2}+\text{const.}.$$ (19) At first this result (i.e. perfect cancellation of the IR singularity in perturbation theory) seems to contradict the general discussion of the non-compact formulation of gauge theories above, where it is shown that transversely separated charges need to be connected by a gauge string in order to cancel the IR divergences. However, this apparent contradiction is resolved by the simple observation the $``$ lattice Hamiltonian for non-compact $`QED_{2+1}`$ coupled to external sources is, technically speaking, just a bunch of coupled shifted harmonic oscillators (the Hamiltonian is Gaussian!). This has two important consequences: First of all, for a shifted harmonic oscillators, the exact ground state energy is obtained already in $`2^{nd}`$ order perturbation theory. Therefore, the calculated ground state energy (18) is the exact one. Secondly, in the language of Fock space operators, the eigenstate of a shifted harmonic oscillator are coherent states, i.e. exponentials of raising operators. Keeping these facts in mind, everything fits together: The coherent states, which are the eigenstates of the the $``$ lattice in the presence of the external source, accomplish the same effect as the exponentials of link fields, namely the act as soliton like operators which are necessary to cancel the small $`q^+`$ divergence of the instantaneous self-interaction. Furthermore, the fact that we are dealing only with shifted harmonic oscillators in $`QED_{2+1}`$ also guarantees that we observed this cancellation already in perturbation theory — even though we did not construct these coherent states explicitly. However, the bad news is that it is difficult to construct a Fock space basis containing exponentials of the gauge fields. <sup>7</sup><sup>7</sup>7Even in $`QED_{2+1}`$, the exact solution contains exponentials of $`a^{}`$ and thus leads to non-normalizable states. Furthermore, when one cannot solve the problem exactly, it is very difficult to achieve this cancellation of IR singularities — unless one works with a basis of gauge invariant states, which is very hard in the non-compact formulation. As usual, in QCD, the situation is worse because the gauge fields themselves carry color-charge. Furthermore, if one tries to maintain gauge invariance then the ansatz $`U_{}=\mathrm{exp}\left(igA_{}\right)`$ leads to $`\frac{1}{g^2}_\mu \mathrm{exp}(igA_{})^\mu \mathrm{exp}(igA_{})=_\mu A_{}^\mu A_{}+\text{“higher orders”}`$ and these higher order terms make it again very difficult to quantize the theory. For these reasons, the non-compact formulation of the $``$ lattice has been abandoned. ### Color-Dielectric Formulation of the $``$ Lattice Naively, one might think that one possibility to introduce $`U_{}SU(N)`$ fields would be to work with linearized general complex matrix fields and to add a potential $`V_{eff}(U_{})`$ that has a minimum for $`U_{}SU(N)`$ (e.g. “Mexican hat” type potential). However, in the Fock expansion (an essential ingredient in the LF Hamiltonian formulation), the fields are usually expanded around the origin (or some other fixed value). Expanding around the origin makes no sense for a Mexican hat shaped potential since the origin corresponds to the false vacuum. On the other hand, expanding around any point along the minimum breaks the manifest global gauge symmetry. Therefore, this idea of adding a (Lagrange multiplier) effective self interaction to enforce the constraint must be abandoned as well. A solution out of this dilemma is to work with blocked (smeared, averaged) degrees of freedom $``$, which are obtained from the original $`U_{}SU(N)`$ by averaging $`U_{}`$s or strings of $`U_{}`$s over some finite volume, e.g. by defining $`=_{av}U`$ dalley . The advantage of this procedure is that the smeared $``$s are no longer subject to the strict $`SU(N)`$ constraint. The blocked theory is still equivalent to the original theory, provided the action for the $``$s contains an effective interaction defined by integrating out the $`U_{}`$s $$\mathrm{exp}\left[V_{eff}()\right]=𝒟U_{}\delta \left(\underset{av}{}U\right)\mathrm{exp}\left[S_{canonical}(U_{})\right].$$ (20) The catch in the whole procedure is that $`V_{eff}()`$ as defined through Eq. (20) can be infinitely complicated and for its exact determination one would have to perform a path integral. However, for approximate calculations, one can always make an ansatz for $`V_{eff}()`$ and there exist various options to determine the parameters appearing in this ansatz. Note that a direct use of Eq. (20) to calculate $`V_{eff}()`$ within the LF framework does not seem to be possible: in order to evaluate the r.h.s. of Eq. (20) one needs to work with link-fields $`U_{}SU(N)`$ and the difficulties in doing this were the main motivation to introduce the color-dielectric formulation in the first place. In the Euclidean, calculating $`V_{eff}`$ seems to be more straightforward, but for using it on the $``$ lattice, the issue arises of translating $`V_{eff}`$ from the Euclidean to the LF. <sup>8</sup><sup>8</sup>8See Ref. mb:sg for a discussion of this issue in the context of scalar field theories. An alternative procedure, based on covariance requirements, appears to be very promising. Due to lack of space, the reader is referred to Refs. d2 ; dalley , where the procedure has been discussed in detail. Acknowledgements: I would like to thank the organizers for the invitation and DFG for financial assistance. It is a pleasure to thank P. Griffin, B. vande Sande and S. Dalley for many enlightening discussions.
no-problem/9908/hep-lat9908055.html
ar5iv
text
# Instanton Effects in Hadron Spectroscopy Revisited ## Abstract We use an optimised clover action to study spectroscopy on an instanton ensemble reconstructed from smoothed Monte Carlo configurations. Due to the better chirality of the clover action, the artificial configurations show a marked difference from the free field behaviour obtained with the Wilson action. They however still fail to reproduce the physics observed on the smoothed configurations. The presence of freely propagating quark modes is found to be responsible for this. According to the instanton liquid model most of the low-energy properties of QCD can be explained by instantons . This however has never been fully tested starting from first principles, in particular on the lattice. In a previous study hadron spectroscopy was performed on instanton ensembles reconstructed from smoothed Monte Carlo configurations . The instanton ensemble was found to exhibit chiral symmetry breaking but the lightest states in the pion and rho channel remained degenerate down to small quark masses; a feature typical of free field theory (no gluons). Now we would like to shed some more light on the origin of this peculiar admixture of “free field like” and “QCD like” properties of the instanton ensemble. We shall compare the following two ensembles of gauge configurations: * Quenched Monte Carlo gauge configurations produced with the fixed point action of at a lattice spacing of $`a=0.144`$fm and then cycled (smoothed with a renormalisation group based procedure) 9 times, up to the point where instantons could be reliably identified. * Artificially constructed instanton configurations reproducing the instanton content (location and size but not the relative orientation in internal space) of the above (for more details on the construction, see Ref. ). We shall refer to the first (smoothed) and the second (instanton) ensemble as SM and IN respectively. The SM ensemble is so smooth that 70% of its action is carried by the instantons (assuming no interaction between them), still it has been shown to contain all the relevant long distance features of QCD. An important feature of the instanton liquid model is the presence of near zero quark eigenmodes that are approximately linear combinations of instanton and antiinstanton zero modes. Their formation is possible only if the zero modes are (at least nearly) degenerate. Only then can small perturbations — e.g. gauge field fluctuations — mix them. This is not the case with the Wilson action. Since it breaks chiral symmetry explicitly, the would be fermion zero modes spread in an interval of the real axis. To check how important the chirality of the fermion action is, we repeated the spectroscopy with the clover action (c<sub>sw</sub>=1.2) optimised to produce “zero modes” in a narrow range around zero . Fig. 1 shows the result for $`m_\pi `$ vs. $`m_\rho `$. This markedly differs from the $`m_\pi =m_\rho `$ free field line (also shown) that was obtained with the Wilson action. This clearly shows the importance of the chiral symmetry of the fermion action. On the other hand, the $`\pi \rho `$ splitting on the IN ensemble still fails to reproduce that on the SM ensemble. In view of the good chirality of the optimised clover action, it is hard to imagine that this can be fully attributed to the remaining explicit chirality breaking of the action. Fig. 2 shows the pion mass sqared vs. the bare quark mass for both ensembles. On the SM ensemble $`m_\pi ^2m_q`$, as expected from PCAC. This is clearly not the case on the IN ensemble, where a best fit to the form $`m_\pi ^\lambda m_q`$ gives $`\lambda =1.5`$ which is between $`\lambda =2`$ and the free field value, $`\lambda =1`$. To obtain a more “microscopic” understanding of the nature of quark propagation in the two ensembles, we also looked at some randomly selected close to zero quark eigenvalues and the corresponding eigenmodes. On a flat gauge field configuration with periodic bondary conditions in all directions, there are $`4N_c`$ trivial (constant) zero modes, this is the number of c-number degrees of freedom corresponding to one fermion flavour. In the presence of antiperiodic boundary condition in the time direction, the lowest eigenmodes are shifted away from zero, to $$\mu _\pm =1\mathrm{cos}\frac{\pi }{N_t}\pm isin\frac{\pi }{N_t}=0.019\pm 0.195i,$$ (1) since in our case $`N_t=16`$. Both eigenvalues are $`4N_c`$-fold degenerate. For purely technical reasons we use antiperiodic boundary condition in the time direction. This separates the lowest free field modes from the lowest topological modes and makes the computation faster. For orientation, in Fig. 3 we plotted the eigenvalues that we used from both ensembles, together with the location of the lowest antiperiodic free field mode. The mixing of the free field modes into the eigenmodes of SM and IN can be characterised by $`P_\pm (\lambda )=P_\pm \psi _\lambda `$, where $`P_\pm `$ is the projection of the normalised eigenmode $`\psi _\lambda `$ onto the eigenspace corresponding to the eigenvalue $`\mu _\pm `$. This quantity is not gauge invariant therefore we work in Lorentz gauge. Any generic fermion mode on a non-flat but locally smooth configuration will have a nonzero projection on the $`\mu _\pm `$ eigenspaces. The question is whether this is just an accidental mixing or the given ensemble really contains close to free field modes. To decide this, a good quantity to look at is $`P(\lambda )=P_+(\lambda )/P_{}(\lambda )`$. If the mixing is accidental, we expect $`P(\lambda )`$ to fluctuate around 1, independently of the corresponding eigenvalue $`\lambda `$. On the other hand, if there are free field like modes on a given configuration then $`P(\lambda )`$ will increase substantially when $`\lambda `$ approaches the value $`\mu _+`$. In Fig. 4 we plot $`P(\lambda )`$ versus the distance of the given eigenvalue from the free field mode $`\mu _+`$. In the SM ensemble $`P(\lambda )`$ fluctuates around 1 everywhere, there is no trace of the free field mode. On the other hand, in the IN ensemble, the modes close to $`\mu _+`$ have a substantially larger projection on the $`\mu _+`$ eigenspace than on the $`\mu _{}`$ subspace. In fact, these modes, close to $`\mu _+`$ have $`P_+(\lambda )1`$, so they are essentially free field modes. A detailed study of the chiral density $`\psi _\lambda ^{}\gamma _5\psi _\lambda `$ reveals that all the $`P(\lambda )1`$ modes look like mixtures of instanton zero modes with the density concentrated in several lumps. This is to be contrasted with the chiral density of the modes with $`P(\lambda )1`$ that spreads roughly homogenously over the whole lattice, as expected of the lowest free field eigenmode. In conclusion, we found that the instanton ensemble reconstructed from smoothed Monte Carlo configurations shows a peculiar mixture of free field and interacting QCD like behaviour. This is a consequence of the presence of both types of quark eigenmodes in this ensemble. The scattering of quarks by instantons is apparently not enough to eliminate free propagation completely. In contrast, we have not found any trace of the free field modes on the smoothed configurations. This is presumably due to some long wavelength fluctuations also responsible for confinement which is absent from the IN ensemble. Finally we note that the density of modes around zero goes as $`V^{1/4}`$ in free field theory, whereas the density of instanton modes in the instanton ensemble is $`V`$, the lattice volume. It is thus concievable that in the absence of confinement, there are always some free field modes which contribute to quark propagation but in the $`V\mathrm{}`$ limit they might be overwhelmed by the instanton modes. It would be interesting — although probably not very cheap — to test this directly on the lattice. The other possibility is that instantons alone might not be enough to reproduce low energy QCD, confinement is also needed. In this case it is confinement that eliminates free quark propagation and makes the hopping of quarks from instanton to instanton the dominant long distance mode of propagation.
no-problem/9908/astro-ph9908039.html
ar5iv
text
# A Two-Dimensional, Self-Consistent Model of Galactic Cosmic Rays in the Heliosphere ## 1 Introduction The past decade has seen a number of advances in spherically-symmetric self-consistent models of the heliosphere that include galactic cosmic rays (GCRs). These have been found to have a complex effect on the heliospheric termination shock. First, the supersonic upstream flow is decelerated by the radial cosmic-ray pressure gradient creating a shock precursor in addition to the gas-pressure-mediated subshock (Drury & Völk (1981)). Second, the shock moves inward closer to the Sun as a result of decreased dynamic pressure of the wind upstream (see Ko, Jokipii, & Webb (1988)). Third, the flow downstream of the shock is modified as well, with the gas pressure being higher just downstream of the shock as a result of momentum conservation (Donohue & Zank (1993)). While, originally, momentum-integrated quantities were used for the particles, recently a more sophisticated approach has been adopted which involves solving the full cosmic-ray transport equation (Parker (1965)). le Roux and Ptuskin (1995) and le Roux and Fichtner (1997) published a series of sophisticated 1-D models that included both galactic and anomalous cosmic rays. 2-dimensional simulations of the cosmic rays as test particles have been available for 2 decades (see, e.g., Jokipii & Kopriva (1979); Jokipii & Davila (1981); Potgieter & Moraal (1985); Haasbroek & Potgieter (1995); Jokipii, Kota, & Merényi (1993)). These showed that drifts in the heliospheric magnetic field play an important role, often more so than diffusion, in transporting the GCRs inside the heliosphere. In particular, for $`A>0`$ (northern polar magnetic field directed outward), GCRs have enhanced access to the interior regions near the pole while at low latitudes their propagation is governed by the outward drift within the current sheet. Since it’s impossible to treat the drifts correctly in 1 dimension, an important part of the physical picture is being left out of one-dimensional models. This paper presents the first results from a 2-dimensional model of the GCR-modified heliosphere. Unlike the recent work of Izmodenov (1997) the structure of the outer heliosphere is greatly simplified here (there is no external interstellar flow and, hence, no heliopause or bow shock), however our description of cosmic-ray transport is much more complete. Our model should only be compared with the upwind side of the heliosphere. The purpose of this work is to illustrate the basic physics of the effects of the GCRs on the flow in the outer heliosphere, not to reproduce in detail actually observed wind flow or particle spectra. While our results are still preliminary, we believe we have discovered an important possible effect of cosmic rays downstream of the termination shock. ## 2 Cosmic-Ray-Modified Solar Wind Our model uses a spherical ($`r\theta `$) coordinate system with symmetry about the polar axis. Another symmetry is about the equatorial plane, which allows us to study the northern hemisphere only; all quantities in the southern hemisphere are their mirror images except the magnetic field components $`B_r`$ and $`B_\varphi `$, which change sign across the neutral sheet. The conservation equations for the wind, neglecting the magnetic force, are $$\frac{\rho }{t}+\frac{(\rho u_i)}{x_i}=0$$ (1) $$\frac{u_i}{t}+u_j\frac{u_i}{x_j}+\frac{1}{\rho }\frac{P_g}{x_i}=\frac{1}{\rho }\frac{P_c}{x_i}$$ (2) $$\frac{}{t}\left(\frac{\rho u^2}{2}+\rho \epsilon \right)+\frac{}{x_i}\left[\rho u_i\left(\frac{u^2}{2}+w\right)\right]=u_i\frac{P_c}{x_i},$$ (3) where $`\rho `$, $`𝐮`$ and $`P_g`$ are the wind mass density, velocity and thermal pressure, respectively. For a polytropic gas, internal energy $`\epsilon =P_g/\rho (\gamma 1)`$ and enthalpy $`w=\gamma P_g/\rho (\gamma 1)`$. $`P_c`$ is the cosmic-ray pressure, determined from the distribution function $`f`$ discussed below. In the supersonic wind, the magnetic force is negligible. We also find the magnetic pressure to be insignificant compared to the thermal and the GCR pressures downstream of the shock. However some of our simulations exhibited regions where the magnetic force could become quite large (see Discussion). Later revisions of our model will include the dynamic effects of the magnetic field on the wind. Parker’s (1965) transport equation is used for the cosmic-ray distribution function $`f`$ $$\frac{f}{t}+u_i\frac{f}{x_i}+v_{d,i}\frac{f}{x_i}\frac{}{x_i}\left(\kappa _{ij}\frac{f}{x_j}\right)=\frac{1}{3}\frac{u_i}{x_i}\frac{f}{\mathrm{ln}p},$$ (4) where $`𝐯_𝐝`$ is the particle drift velocity and $`\kappa `$ is the (symmetric) diffusion tensor. For protons, the relevant drift velocity components can be written as $$v_{d,r}=\frac{pcv}{3e}\frac{1}{r\mathrm{sin}\theta }\frac{}{\theta }\left(\mathrm{sin}\theta \frac{B_\varphi }{B^2}\right),v_{d,\theta }=\frac{pcv}{3e}\frac{1}{r}\frac{}{r}\left(r\frac{B_\varphi }{B^2}\right),$$ (5) where $`p`$ and $`v`$ are particle’s momentum and speed, respectively and $`e`$ is the elementary charge. For diffusion we adopted a commonly-used form (e.g., Jokipii & Davila (1981), Jokipii, et al. 1993): $$\kappa _{ij}=\kappa _{}\delta _{ij}+\frac{\left(\kappa _{}\kappa _{}\right)B_iB_j}{B^2},\kappa _{}=\kappa _0P^{1/2}\frac{v}{c}\frac{B_0}{|B|},$$ (6) where $`P`$ is the particle’s rigidity and $`B_0`$, the magnetic field intensity at 1 AU distance in the ecliptic plane. In our simulations we took $`\kappa _0=1.510^{22}\text{cm}^2\text{s}^1`$ and $`\kappa _{}=0.05\kappa _{}`$. We used a uniform, latitude-independent 400 km/s solar wind as the initial condition. The external gas pressure was chosen such that the shock was initially a sphere located at approximately 80 AU distance from the Sun (similar to le Roux & Fichtner (1997)). Our model can readily handle non-spherical shocks. However, this simple initial geometry helps us to isolate purely cosmic-ray-induced behavior vs. effects caused by latitudinal dependence in the wind dynamic pressure. The solar wind equations were solved using the second order Godunov-type numerical scheme (Colella (1990)) and the HLLE Riemann solver of Einfeldt, et al. (1991). The cosmic-ray equation (4) was solved using a second order ADI split scheme based on the scheme of McKee and Mitchell (1970). ## 3 Magnetic Field and Cosmic-Ray Transport The current version of our model neglects the magnetic force on the wind and Faraday’s law can be solved independently. For the azimuthal component we have $$\frac{B_\varphi }{t}=\frac{1}{r}\frac{\left(ru_rB_\varphi \right)}{r}\frac{1}{r}\frac{\left(u_\theta B_\varphi \right)}{\theta }.$$ (7) The other two components of the field are usually much smaller, except at small radii, but they still are computed for consistency using the steady-state version of the Faraday’s law $`u_rB_\theta =u_\theta B_r`$. This approach is justified because of the long evolutionary time scales of the system. Combined with the Maxwell’s equation $`𝐁=0`$, the above expression can be solved for the radial field component. To account for the fluctuating component of the magnetic field observed in the polar regions of the solar wind we included the modified latitudinal field (Jokipii & Kota (1989), Jokipii, et al. (1995)), $`B_\theta ^m`$ in the overall magnetic field structure. Consider the motion of the magnetic field footpoints due to supergranulation or some other random process, in the $`\theta `$ direction. If we assume a simple, periodic supergranulation motion, then $`B_\theta ^m=B_{\theta 0}^mg(r,t)`$ where $`g`$ is a rapidly-oscillating function. The wavelengths and periods of the fluctuations are expected to be much smaller than the spacial and time scales of the system. By averaging the $`\theta `$ component of the Faraday’s law, multiplied by $`g`$, it may be shown that $$\frac{B_{\theta 0}^m}{t}=\frac{1}{r}\frac{\left(ru_rB_{\theta 0}^m\right)}{r},$$ (8) where we used $`g_{r,t}=gg/r_{r,t}=gg/t_{r,t}=0`$ and $`gg_{r,t}0`$. A modified field computed in this fashion enters into $`B^2`$ which results in transport coefficients that are smaller by several orders of magnitude at the pole, according to equations (5) and (6). Since the magnetic field changes polarity across the equatorial current sheet, the drift velocity contains a delta-function. We employed two different approaches in dealing with the current sheet. In the first we treat the drift velocity as a boundary conditions as in Jokipii & Kopriva (1979). We also carried out calculations for the wide current sheet model of Potgieter & Moraal (1985). The two sets of results were consistent. All results presented here are for a current sheet with a half-width of $`8^0`$. ## 4 Model Results We have performed a number of simulations with different resolutions and equatorial boundary conditions. For the results presented here the spatial resolution was 1 AU in the radial direction and $`0.5^0`$ in latitude. At the outer boundary of the simulation, the gas pressure was taken to be uniform at 1.1 $`\text{eV cm}^3`$ and the GCR pressure was 0.4 $`\text{eV cm}^3`$. The cosmic-ray spectrum at the far boundary was the same as in Jokipii & Kopriva (1979). The inner, cosmic-ray-absorbing, boundary was placed at 5 AU to avoid numerical errors due to insufficient resolution in the region of the rapid wind expansion. Results presented here are not for a steady-state configuration. In fact we have never achieved a real steady state except in case of the small cosmic-ray content. However, the rate of change of most parameters slows down after a few years. After a sufficiently long period of time (of the order of 10 years) we begin to see an inflow at the outer boundary at the pole for $`A>0`$ or the ecliptic, for $`A<0`$, which cannot be handled properly by the present boundary conditions. Because the magnetic field reversal occurs every 11 years, we present here results for the state 8 years from the start of the simulation. Figures 1a and 1b show the radial solar-wind speed contours and streamlines for $`A>0`$ and $`A<0`$ respectively. We can see a significant change in the flow pattern downstream of the shock. The wind turns away from the pole and towards the equator in the case $`A>0`$ while the opposite is true for $`A<0`$. Since the subsonic flow tends to be incompressible, its density varies little at mid-latitudes, all changes are concentrated near the pole and the equatorial neutral sheet. Note that the shock remains essentially spherical, even though there is clearly a strong latitudinal dependence of the various parameters. The subshock compression ratio varies slightly between 3.8 at the equator and 4.2 at the pole (for $`A>0`$). Figures 2a and 2b show the intensity contours for 1.6 GeV protons, and Figure 3 plots the same quantity for 72 MeV protons, both figures also show test-particle results for comparison (dashed lines). We find that for $`A>0`$, the cosmic-ray distribution is only slightly affected by the changing conditions in the solar wind. Here radial transport coefficients become slightly smaller both near the equator (because of compression, $`B_\varphi `$ is increased) and in the polar region (due to enhanced $`B_\theta ^m`$). The small change in intensity agrees with results of Jokipii & Davila (1981) who found that the particles are not strongly affected by relatively small changes in transport coefficients. Most of the difference we see is probably due to the changed position of the shock. However, for $`A<0`$ there is a larger difference in particle intensity at high latitudes. This is a result of higher degree of wind modification near the pole than in the $`A>0`$ case. Figure 4 shows equatorial and polar thermal and GCR pressures and the radial wind velocities for the present epoch ($`A>0`$). The termination shock has moved inward between 3 and 4 AU. A wide shallow cosmic-ray shock precursor is visible in the supersonic wind. Notice that the wind is faster at the equator as a result of compression and slower near the pole due to expansion. Equatorial gas pressure build up is due to both compression and conservation of momentum downstream of the shock (the sum, $`P_g+P_c`$, tends to remain nearly a constant). The pressure peak just downstream of the shock at the equator corresponds to a region of large $`u_\theta `$ at slightly higher latitude and seem to balance the dynamic pressure of the equatorward flow. However, it could possibly be an artifact of the solution. The largest dynamic pressure, $`\rho u_\theta ^2/2`$, associated with the latitudinal flow is about 0.013 $`\text{eV cm}^3`$ while the pressure peak is 0.08 $`\text{eV cm}^3`$ above the average gas pressure in this region. For $`A<0`$, wind velocity and thermal pressure profiles at the pole look similar to those at the equator for $`A>0`$ and the peak is present, too. ## 5 Discussion With a series of tests we have been able to verify that the $`\theta `$ gradients of the cosmic-ray pressure are responsible for the solar wind flow configuration we observed. We were able to produce a general flow structure similar to that shown in Figure 1 by artificially placing a constant in time cosmic-ray pressure gradient that varied in latitude only. In turn, large lateral gradients are due to difference in radial drift velocity between the pole and the equator, in agreement with Jokipii & Davila (1981). Because latitudinal drifts and diffusion tend to smooth out these gradients the wind will be less affected. The fact that the shock is still spherical remains unexplained, it could be due to our pressure boundary conditions that are spherically symmetric. We have found that the system continues to evolve on time scales comparable to the 11-year solar cycle. We expect the heliospheric configuration to oscillate between states similar to those shown on Figures 1a-1b and 2a-2b. We have not seen the system reach a steady state under stationary boundary conditions when the cosmic-ray pressure is comparable to the thermal pressure at the external boundary. Notice a large positive radial gradient in $`P_g`$ in the polar region in the $`A>0`$ case (Figure 4). There the subsonic wind is propagating away from the Sun against this gradient and can be quite easily turned backward. Such an inflow would require specifying two extra conditions at the outer boundary. Should an inflow occur, the magnetic field is dragged along the lines separating the external flow and the solar wind. Because the field offers no resistance, both $`B_\theta `$ and $`B_\theta ^m`$ may become quite large. To see if this effect has physical significance we plan to add magnetic force to the solar wind equations. In addition, anomalous cosmic rays should contribute significantly to the dynamics of the system; we plan to include them in a later publication. This work was supported, in part, by NASA under grants NAG5-6620, NAG5-7793 and by the NSF under grant ATM9616547.
no-problem/9908/hep-lat9908027.html
ar5iv
text
# Magnetic Monopoles, Center Vortices and Topology of Gauge Fields ## 1 Introduction Recent lattice calculations have given strong evidence for two confinement scenarios: 1. the dual Meissner effect , which is based on a condensate of magnetic monopoles in the QCD vacuum and 2. the center vortex picture , where the vacuum consists of a condensate of magnetic flux tubes which are closed due to the Bianchi identity. There are also lattice calculations which indicate that the spontaneous breaking of chiral symmetry, which is related to the topology of gauge fields, is caused by these objects, i.e. by either magnetic monopoles or center vortices . In this talk we would like to discuss the topological properties of magnetic monopoles and center vortices. We will first show that in Polyakov gauge the magnetic monopoles completely account for the non-trivial topology of the gauge fields. Subsequently, we will extend the notion of center vortices to the continuum. We will present the continuum analog of the maximum center gauge fixing and the Pontryagin index of center vortices. ## 2 Magnetic monopoles and topology The magnetic monopoles arise in Yang-Mills-Theories in the so called Abelian gauges . For the study of these monopoles in the continuum theory the Polyakov gauge is particularly convenient. In this gauge one diagonalizes the Polyakov loop $`\mathrm{\Omega }(\stackrel{}{x})=Pe^{_0^T𝑑x_0A_0(x_0,\stackrel{}{x})}=V^+\omega V`$ (1) which fixes $`VSU(N)/U(1)^{N1}`$ i.e. the coset part of the gauge group. Magnetic monopoles arise as defects of the gauge fixing which occur when at isolated points in space $`\stackrel{}{x}_i`$ the Polyakov loop becomes a center element $`\mathrm{\Omega }(\stackrel{}{x}_i)=(1)^{n_i},n_i:\text{integer}`$ (2) The field $`A^V=VAV^++VV^+`$ develops then magnetic monopoles. These monopoles have topologically quantized magnetic charge given by the winding number $`m[V]\mathrm{\Pi }_2(SU(2)/U(1))`$ (3) of the mapping $`V(\stackrel{}{x})`$ from a sphere $`S_2`$ around the magnetic monopole into the coset $`SU(2)/U(1)`$ of the gauge group. In the Polyakov gauge the Pontryagin index can be exactly expressed in terms of magnetic charges. If we assume a compact space-time manifold and that there are only point-like defects of gauge fixing, i.e magnetic monopoles are the only magnetically charged objects arising after gauge fixing, the Pontryagin index is given by $`\nu =\mathrm{\Sigma }_in_im_i`$ (4) The summation runs here over all magnetic monopoles with $`m_i`$ being the magnetic charge of the monopole defined by equation (3) and the integer $`n_i`$ is defined by the value of the Polyakov-loop at the monopole position (2). This relation shows that the magnetic monopoles completely account for the non-trivial topology of gauge fields, at least in the Polyakov gauge. Unfortunately, in other Abelian gauges like maximum Abelian gauge, such a simple relation between Pontryagin index and topological charges is not yet known. However, in the maximum Abelian gauge correlations between instantons and monopoles has been found, in both analytical and lattice studies . ## 3 Center vortices in the continuum In order to extend the notion of center vortices to the continuum theory let us, undertake a detour through the lattice by putting a given smooth gauge field $`A_\mu (x)`$ on a fine lattice in the standard fashion by introducing the link variables $`U_\mu (x)=exp(aA_\mu (x))`$. For initially smooth gauge fields $`A_\mu (x)`$ and for sufficiently fine lattices all link variables will be close to unity, which represents the ”north pole” of the SU(2) group manifold $`S_3`$. However, in the process of gauge fixing, gauge transformations are performed which transform some of the links from the northern to the southern hemisphere. In fact, we can separate each gauge transformation into a transformation which merely switches a link from the northern to the southern hemisphere (or vice versa) and one which rotates link variables inside either the northern or the southern hemisphere but does not switch hemispheres. In the maximum center gauge condition $$\underset{x,\mu }{}(T_rU_\mu (x))^2max,$$ (5) which is obviously insensitive to center gauge transformations, one exploits gauge transformations to rotate a link variable as close as possible to a center element. Once the maximum center gauge has been implemented, center projection implies to replace all links by their closest center element. One obtains then a $`Z(2)`$ lattice which contains $`D1`$ dimensional hypersurfaces $`\mathrm{\Sigma }`$ on which all links take a non-trivial center element that is $`U=1`$ in the case of $`SU(2)`$. The $`D2`$ dimensional boundaries $`\mathrm{\Sigma }`$ of the hypersurfaces $`\mathrm{\Sigma }`$ represent the center vortices, which, when non-trivially linked to a Wilson loop, yield a center element for the latter. A carefully analysis shows that the continuum analogies of the center vortices are defined by the gauge potential , $`𝒜_\mu (x,\mathrm{\Sigma })=E{\displaystyle _\mathrm{\Sigma }}d^{D1}\stackrel{~}{\sigma }_\mu \delta ^D(x\overline{x}(\sigma ))`$ (6) where $`d^{D1}\stackrel{~}{\sigma }_\mu `$ is the dual of the $`D1`$ dimensional volume element. Furthermore, the quantity $`E=E_aH_a`$ with $`H_a`$ being the generators of the Cartan algebra represents (up to a factor of $`2\pi `$) the so called co-weights which satisfy $`exp(E)=ZZ(N)`$. Due to this fact the Wilson-loop calculated from the gauge potential (6) becomes, $`W[𝒜](C)=\mathrm{exp}({\displaystyle _C}𝒜)=Z^{I(C,\mathrm{\Sigma })}`$ (7) where $`I(C,\mathrm{\Sigma })`$ is the intersection number between the Wilson-loop $`C`$ and the hypersurface $`\mathrm{\Sigma }`$. The representation, (6), is referred to as ideal center vortex. One should emphasize that the hypersurface $`\mathrm{\Sigma }`$ can be arbitrarily deformed by a center gauge transformation keeping, however, its boundary $`\mathrm{\Sigma }`$, i.e. the position of the center vortex, fixed. Thus for fixed $`\mathrm{\Sigma }`$ the dependence of the gauge potential (6) on the hypersurface itself is a gauge artifact. The dependence on the hypersurface $`\mathrm{\Sigma }`$ can be removed by performing the gauge transformation $`\phi (x,\mathrm{\Sigma })=exp(E\mathrm{\Omega }(x,\mathrm{\Sigma }))`$ (8) where $`\mathrm{\Omega }(x,\mathrm{\Sigma })`$ is the solid angle subtended by the hypersurface $`\mathrm{\Sigma }`$ as seen from the point $`x`$. One finds then $`𝒜_\mu (x,\mathrm{\Sigma })=\phi (x,\mathrm{\Sigma })_\mu \phi ^+(x,\mathrm{\Sigma })+a_\mu (x,\mathrm{\Sigma })`$ (9) where $`a_\mu (x,\mathrm{\Sigma })=E{\displaystyle _\mathrm{\Sigma }}d^{D2}\stackrel{~}{\sigma }_{\mu \nu }_\nu D(x\overline{x}(\sigma ))`$ (10) depends only on the vortex position $`\mathrm{\Sigma }`$ and is referred to as ”thin vortex”. Here $`D(x\overline{x}(\sigma ))`$ represents the Green function of the $`D`$ dimensional Laplacian. In fact, one can show that the thin vortex represents the transversal part of the ideal vortex $`a_\mu (x,\mathrm{\Sigma })=P_{\mu \nu }𝒜_\nu (x,\mathrm{\Sigma })`$ where $`P_{\mu \nu }=\delta _{\mu \nu }\frac{_\mu _\nu }{^2}`$ is the usual transversal projector. A careful and lengthy analysis yields then the conclusion that the continuum analog of the maximum center gauge fixing is defined by $`\underset{\mathrm{\Sigma }}{\mathrm{min}}\underset{g}{\mathrm{min}}{\displaystyle \text{Tr}(A^ga(\mathrm{\Sigma }))^2}`$ (11) where the minimization is performed with respect to all (continuum) gauge transformations $`gSU(2)/Z(2)`$ (which represents per se coset gauge transformations) and with respect to all vortex surfaces $`\mathrm{\Sigma }`$. For fixed thin center vortex field configuration $`a(\mathrm{\Sigma })`$ minimization with respect to the continuum gauge transformation g yields the background gauge condition $`[_\mu +a_\mu (\mathrm{\Sigma }),A_\mu ]=0`$ (12) where the thin vortex field $`a_\mu (x,\mathrm{\Sigma })`$ figures as background gauge field. One should emphasize, however, that the background field has to be dynamically determined for each given gauge field $`A_\mu (x)`$ and thus depends on the latter. Obviously in the absence of a vortex structure in a considered gauge field $`A_\mu (x)`$ the background gauge condition reduces to the Lorentz gauge $`_\mu A_\mu =0`$. ## 4 Topology of Center vortices Once the center vortex configurations in the continuum are at our disposal, it is straightforward to calculate their Pontryagin index. In the continuum formulation where center vortices live in the Abelian subgroup by construction the direction of the magnetic flux of the vortices is fully kept. The explicit calculation shows that the Pontryagin index $`\nu `$ of the center vortices is given by $`\nu =\frac{1}{4}I(\mathrm{\Sigma },\mathrm{\Sigma })`$ where $`I(\mathrm{\Sigma },\mathrm{\Sigma })`$ represents the self-intersection number of the closed vortex sheet $`\mathrm{\Sigma }`$. A careful analysis shows that for closed orientable surfaces the self intersection number vanishes. In order to have a non-trivial Pontryagin index the vortex surfaces have to be not globally orientable, i.e., they have to consist of orientable pieces. One can further show that at the border between orientable vortex patches magnetic monopole currents flow. It is these monopole currents which make the vortex sheet non orientable since they change the orientation of the magnetic flux. Thus we obtain that even for the center vortices the non-trivial topology is generated by magnetic monopole currents flowing on the vortex sheets. This is consistent with our finding in the Polyakov gauge where the Pontryagin index was exclusively expressed in terms of magnetic monopoles . By implementing the maximum center gauge condition in the continuum one can derive, in an approximate fashion, an effective vortex theory, where the vortex action can be calculated in a gradient expansion. The leading order gives the Nambu-Goto action while in higher orders curvature terms appear. A model based on such an effective vortex action, in fact, reproduces the gross features of the center vortex picture found in numerical lattice simulations. ## Acknowledgment: This work was supported in part by the DFG-Re 856/4-1 and DFG-En 415/1-1.
no-problem/9908/gr-qc9908016.html
ar5iv
text
# Symmetry-seeking spacetime coordinates ## I Introduction One of the fundamental problems in numerically constructing spacetimes without symmetries (3D numerical relativity) is the choice of suitable coordinates. The task is not just to find a good coordinate system for a known spacetime, but the spacetime is not actually known when the calculation begins, so that one has to construct the coordinate system along with the spacetime. In principle, once one has obtained the spacetime in some coordinate system, one can go back over it and paint on a better coordinate system. In practice, the initial attempt must not be too bad, or the coordinate system will break down altogether before one has evolved very far. One begins the numerical construction of a spacetime with Cauchy data (three-metric $`g_{ab}`$, extrinsic curvature $`K_{ab}`$, and perhaps suitable matter data) obeying the Hamiltonian and momentum constraints, in hand. One then has to make a choice of the lapse $`\alpha `$ and shift $`\beta ^a`$ based on this information in order to start constructing the spacetime along with the coordinate system. For the purpose of this paper, a coordinate condition is a prescription that maps Cauchy data on a Cauchy surface to lapse and shift fields on that Cauchy surface. One then uses the same prescription again on each time slice. We call such coordinate conditions local in time. Typically they are elliptic (in space), and one specifies suitable boundary conditions at the outer (large radius) and inner (black hole excision) boundaries of the numerical domain. In a generalization of local-in-time coordinate conditions, one may use a prescription for the first or second time derivatives of the lapse and/or shift, obtaining a heat equation or wave equation-like prescription. Such evolution equations for the lapse and shift require much less computational work than solving an elliptic equation at each time step. Coordinate conditions for the purpose of our paper are then elliptic, parabolic or wave equations for the lapse and shift in terms of given Cauchy data on a slice. What are the characteristics of good coordinate conditions? Here we propose the following desirable criterion: If the spacetime is in fact stationary, the coordinate condition should make the metric coefficients explicitly time-independent, independently of how the initial slice is embedded in the spacetime. If there is an approximate timelike Killing vector, the metric should be approximately time-independent. If there is no Killing vector, the natural extension of our proposal is to minimize some measure of the rate of change of the metric with time. Consider the inspiral phase of a binary system. There is an approximate timelike Killing vector here, in the sense that all metric coefficients can be made to evolve on the inspiral time scale instead of the much shorter orbital time scale. Clearly, this defines a corotating coordinate system. If the spacetime has an approximate periodicity (discrete isometry connecting timelike related points), as in the case of a highly elliptic inspiral, the metric should become approximately periodic in time through our coordinate condition. A coordinate system adapted to the presence of a timelike Killing vector, as we have just described it, is far from unique. It is easy to see that the family of such coordinate systems is equivalent to the family of smooth spacelike slices containing one arbitrary but fixed spacetime point, together with spatial coordinates on the slice: the coordinate system on the spacetime is obtained by Lie-dragging the slice and its coordinates along the Killing vector. One may know that the final state of an evolution (for example a binary black hole system) is a Kerr solution, but this does not mean that the metric coefficients approach the Kerr metric in any of the standard coordinate systems even as they become asymptotically time-independent. We require only that our coordinate condition evolves any given initial slice along the timelike Killing vector, not that it “straightens out” the slice. A “straightest” slice in this spacetime could be defined as one that minimizes for example $`{}_{}{}^{(3)}R_{}^{2}𝑑V`$, but we shall not try to minimize this at the same time. A mathematically similar problem to that of finding good coordinates in numerical relativity arises in the theory of critical phenomena in gravitational collapse. The entire rich phenomenology of critical collapse can be explained and quantitatively predicted using ideas taken from renormalisation group theory and the theory of dynamical systems . Implicit in this theoretical framework is the concept of a coordinate prescription (local in time) such that the metric is “time”-independent if the Cauchy data generate a spacetime that is homothetic, or continuously self-similar (CSS). We have set “time” in quotation marks here, as $`/t`$ is not everywhere timelike in a homothetic spacetime. In this context we are only interested in a prescription that is local in time (in the sense of not containing time derivatives of the lapse and shift). Such a prescription turns general relativity into a dynamical system, with CSS spacetimes corresponding to fixed points. If the spacetime is discretely self-similar (DSS), we require our prescription to make the metric periodic in “time”, with a DSS spacetime corresponding to a limit cycle. (The use of maximal slicing in critical gravitational collapse does part of the job of turning general relativity into the appropriate sort of dynamical system. However, one is still left with the question of what spatial coordinates to use). Again, such a coordinate system is far from unique. In the neighborhood of any data set generating a CSS or DSS spacetime one can define such a coordinate prescription perturbatively. This perturbative definition has been enough in order to calculate critical exponents to several significant digits. A coordinate condition applicable to all of superspace, with the property that the fixed points of the resulting dynamical system are precisely CSS spacetimes and the limit cycles the DSS spacetimes, would define a renormalisation group for general relativity. Studying its properties would help us in understanding the general dynamics of general relativity and issues relating to cosmic censorship. In this paper, we study the two related cases in a common framework. In section II, we discuss a number of candidate coordinate conditions in the absence of symmetries (other than stationarity or homothety). Our basic strategy is to write down the condition that the metric is (conformally) time-independent in a 3+1 split, and then choose four equations out of that set of equations to solve for the lapse and shift or their time derivatives. This will guarantee that our criterion is obeyed. If a given prescription obeying this criterion is otherwise a good coordinate condition must be determined empirically. In section III, we specialize the general proposals to spherical symmetry, and we discuss additional proposals that are specific to spherical symmetry. In section IV, we test some of these proposals by evolving a slice in a given spherically symmetric spacetime. (This procedure is simpler and more stable than evolving the Cauchy data themselves.) We use several spacetimes to test the stationary case, the CSS case, and the DSS case. ## II Coordinate choices ### A 3+1 split of the Killing and homothetic equation We begin with a homothetic Killing vector $`X^a`$. That is $$_Xg_{ab}=_aX_b+_bX_a=2\sigma g_{ab},$$ (1) where $`\sigma `$ is a constant. For $`\sigma =0`$, $`X^a`$ is a Killing vector. For $`\sigma 0`$ a constant we can, without loss of generality, choose $`\sigma =1`$ by rescaling $`X^a`$. In the following $`\sigma `$ only takes the values $`0`$ and $`1`$. We perform the usual 3+1 split of the Einstein equation with $`X^a`$ as the evolution vector field. That is, spacetime is foliated by surfaces $`\mathrm{\Sigma }(t)`$ with $`X^a=(/t)^a`$. Each surface has a unit normal vector $`n^a`$, intrinsic metric $`h_{ab}=g_{ab}+n_an_b`$ with associated derivative operator $`D_a`$ and extrinsic curvature $`K_{ab}=(1/2)_nh_{ab}`$. (There is no standard convention for the sign of $`K_{ab}`$. Here we use the sign prevalent in numerical relativity.) We decompose $`X^a`$ as $`\alpha n^a+\beta ^a`$ with $`\beta ^an_a=0`$. We want to know how the quantities ($`\alpha ,\beta ^a,h_{ab},K_{ab}`$) evolve with time. We will get one answer from the homothetic Killing equation and another answer from the usual ADM equations. The condition that these two answers agree then gives conditions that these quantities must satisfy in a spacetime with a homothetic Killing vector. . We begin by deriving evolution equations from the homothetic Killing equation. From the decomposition of $`X^a`$, it follows that $`n_a=\alpha _at`$ and therefore that $$_Xn_a=\alpha ^1(_X\alpha )n_a.$$ (2) Since $`_Xn_a`$ is a scalar times $`n_a`$, it follows from equation (1) that $$_Xn_a=\sigma n_a,$$ (3) and therefore that $$_X\alpha =\sigma \alpha .$$ (4) Then using $`0=_XX^a=0`$ we find $$_X\beta ^a=0.$$ (5) It follows from equations (1) and (3) that $$_Xh_{ab}=2\sigma h_{ab}.$$ (6) We then find $$_XK_{ab}=\frac{1}{2}_X_nh_{ab}=\frac{1}{2}\left(_n_Xh_{ab}+_{[X,n]}h_{ab}\right),$$ (7) and therefore $$_XK_{ab}=\sigma K_{ab}.$$ (8) ### B Solving for the lapse and shift locally in time We begin by looking at those components of the homothetic Killing equation that contain time derivatives of the lapse and shift. With the shorthand $`_XT=\dot{T}`$ for any tensor $`T`$, these are $$\dot{\alpha }=\sigma \alpha ,\dot{\beta }^a=0.$$ (9) Given a good initial guess, we could consider these as evolution equations for the lapse and shift, but this is unlikely to make for a stable scheme when the symmetry is only approximate and/or the initial guess is not perfect. Furthermore, it leaves us with the problem of finding a good initial value for the lapse and shift. Conversely, if we have found a lapse and shift on the initial time slice by solving some elliptic equations, the only reason not to use the same prescription on each time slice would be numerical effort. \[If numerical effort is a concern, one could in effect evolve the lapse and shift by (9) for a few time steps, before using the expensive prescription again.\] We now concentrate on the other components of the homothetic Killing equations. The ADM evolution equations are $`\dot{h}_{ab}`$ $``$ $`D_a\beta _b+D_b\beta _a2\alpha K_{ab},`$ (10) $`\dot{K}_{ab}`$ $``$ $`\beta ^cD_cK_{ab}+K_{ac}D_b\beta ^c+K_{bc}D_a\beta ^cD_aD_b\alpha +\alpha \left({}_{}{}^{(3)}R_{ab}^{}+KK_{ab}2K_{ac}K_{b}^{}{}_{}{}^{c}\tau _{ab}{\displaystyle \frac{1}{2}}(\rho \tau )h_{ab}\right),`$ (11) for any vector field $`X^a`$. Here $`\rho G_{ab}n^an^b`$, $`\tau _{ab}h_{a}^{}{}_{}{}^{c}h_{b}^{}{}_{}{}^{d}G_{ab}`$, and $`J_ah_{a}^{}{}_{}{}^{c}n^bG_{cb}`$. Note that in the ADM evolution equations, $`\alpha `$ and $`\beta ^a`$ can be freely specified, so no new information is obtained by combining the ADM equations with the equations (9). But if we substitute the ADM equations into the components $$\dot{h}_{ab}=2\sigma h_{ab}$$ (12) $$\dot{K}_{ab}=\sigma K_{ab}.$$ (13) of the homothetic Killing equation, we obtain two symmetric tensor equations which link the lapse and shift, and their spatial derivatives, to the Cauchy data, and their spatial derivatives, all within a single Cauchy surface. Our task can now be formulated as follows. Out of these 6+6 equations we want to obtain 3+1 equations that can be solved for the lapse and shift, for given generic Cauchy data. The remaining 8 equations remain as consistency conditions which are obeyed if and only if the Cauchy data actually evolve into a spacetime with a (conformal) Killing vector. In other words, given Cauchy data on a slice, we want to have a prescription to calculate a lapse and shift such that one obtains a reasonable coordinate system on the spacetime evolved from “reasonable” generic data, and coordinates adapted to the symmetry if there is a timelike Killing vector or homothetic vector. Clearly, there is an infinite range of possibilities. In a situation without symmetries, our task is to make one 3-vector (the equation for the shift) and one scalar (the equation for the lapse) out of two symmetric 3-tensors. It is important to note that the equations for the lapse and shift will generally be coupled, so that the same equation for, say, the shift can have a different character when coupled to a different equation for the lapse. Some or all of the resulting equations can be purely algebraic. Alternatively, they can involve first or higher spatial derivatives of the lapse and shift. An elliptic equation or equations would be particularly appealing. Any prescription obtained for the general case can be reduced to spherical symmetry, and we shall do this for testing. Additional prescriptions that explicitly use the spherical symmetry do not generalize back to the case without symmetries. One can obtain a 3-vector of equations from a 3-tensor by taking a divergence, for example $$D^a\left(\dot{h}_{ab}2\sigma h_{ab}\right)=D^a\dot{h}_{ab}=0.$$ (14) This equation can be derived from varying the action $$d^3x\sqrt{h}h^{ac}h^{bd}(\dot{h}_{ab}2\sigma h_{ab})(\dot{h}_{cd}2\sigma h_{cd})$$ (15) with respect to the shift. As $`\sigma `$ drops out of (14), it could also have been derived from varying the action $$d^3x\sqrt{h}h^{ac}h^{bd}\dot{h}_{ab}\dot{h}_{cd}.$$ (16) Condition (14) was first suggested as a shift condition by Smarr and York under the name minimal strain shift. As it is derived from a variational principle, one may hope that it gives good spatial coordinates also in the absence of an exact symmetry. Written out explicitly, minimal strain shift is $$D_bD^b\beta ^a+D_bD^a\beta ^b2D_b(\alpha K^{ab})=0.$$ (17) Taking the divergence of (13), we obtain the 3-vector equation $$D^a\left(\dot{K}_{ab}\sigma K_{ab}\right)=0.$$ (18) At first sight this is less appealing than (14) because it contains third spatial derivatives of the lapse. In order to obtain a scalar equation for the lapse, we can take the trace of (12); $$h^{ab}\left(\dot{h}_{ab}2\sigma h_{ab}\right)=0$$ (19) is equivalent to $$\dot{\sqrt{h}}3\sigma \sqrt{h}=0,$$ (20) that is, it determines the scaling of the 3-volume element. ($`h`$ is the determinant of $`h_{ab}`$.) Written out in terms of $`\alpha `$ and $`\beta ^a`$, this equation is $$D_a\beta ^a\alpha K3\sigma =0.$$ (21) This does not seem to be in general a good equation for the lapse (since it is ill defined where $`K`$ vanishes), nor (by virtue of being a scalar) for the shift. The equation resulting from instead contracting (12) with $`K_{ab}`$, $$K^{ab}\left(\dot{h}_{ab}2\sigma h_{ab}\right)=0$$ (22) can be derived from varying the functional (15) with respect to the lapse. We shall refer to Eq. (22) on its own as the conformal strain lapse for $`\sigma =1`$, or the minimal strain lapse for $`\sigma =0`$. It is an algebraic equation for the lapse, $$\alpha =\frac{K^{ab}D_a\beta _b\sigma K}{K^{ab}K_{ab}},$$ (23) which is linear in the shift. The fact that the lapse may be zero, negative, or even blow up (where $`K_{ab}=0`$), is worrisome. Clearly this prescription cannot be used on all initial data. For the Killing case, $`\sigma =0`$, Brady, Creighton and Thorne (BCT) have recently proposed using the minimal strain lapse and minimal strain shift together, as both are derived from the same variational principle (16). If one substitutes the formal minimal strain lapse (23) into the minimal strain shift equation (17), it is not clear if the resulting differential equation for the shift alone is elliptic. We shall refer to the combination of minimal strain lapse and minimal strain shift, in the Killing case $`\sigma =0`$, as the BCT gauge, and for any $`\sigma `$ as the generalized BCT gauge. The equation $$h^{ab}\left(\dot{K}_{ab}\sigma K_{ab}\right)=0$$ (24) is an elliptic equation for the lapse, and seems not to have been considered before. We shall call it the trace-Kdot lapse. Adding to this the minimal strain lapse equation, however, we obtain $$0=h^{ab}\left(\dot{K}_{ab}\sigma K_{ab}\right)K^{ab}\left(\dot{h}_{ab}2\sigma h_{ab}\right)=\dot{K}+\sigma K.$$ (25) For $`\sigma =0`$, this is known as constant (extrinsic) curvature slicing. If $`K=0`$ initially, and therefore at all times, we have maximal slicing. For $`\sigma =1`$ we could speak of exponential curvature slicing. Written out in terms of the Cauchy data, the equation is $$D_aD^a\alpha +\left[{}_{}{}^{(3)}R+K^2+\frac{1}{2}(\tau 3\rho )\right]\alpha +\beta ^aD_aK+\sigma K=0.$$ (26) Note that we do not assume that $`K`$ is constant within each slice. We shall refer to the combination of equations (17) and (26) as the generalized Smarr-York (SY) gauge. For an initial $`K=0`$ slice, the generalized SY gauge reduces to maximal slicing with minimal strain. The differential equations for lapse and shift must be supplemented with boundary conditions in order to yield a solution. For an approximate timelike Killing field in an asymptotically flat spacetime, the condition $`\alpha 1`$ and $`\beta ^a0`$ at infinity should be reasonable. For the case of binary black hole inspiral, BCT have proposed a boundary condition to use with their equation. It is less clear what would be a reasonable boundary condition to use for the case of critical gravitational collapse. ## III Spherical symmetry ### A Reduction to spherical symmetry In spherical symmetry, the metric has the form $$ds^2=(\alpha ^2+\beta ^r\beta _r)dt^2+2\beta _rdrdt+g_{rr}dr^2+g_{\theta \theta }(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2)$$ (27) where $`\alpha ,\beta ^r,g_{rr}`$ and $`g_{\theta \theta }`$ are functions of $`r`$ and $`t`$, and where $`\beta _r=g_{rr}\beta ^r`$. Equations (12) and (13) now have only two independent components each. The $`rr`$ and $`\theta \theta `$ components of equation (12) are respectively $`{\displaystyle \frac{1}{2}}\beta ^r{\displaystyle \frac{dg_{rr}}{dr}}+g_{rr}{\displaystyle \frac{d\beta ^r}{dr}}\alpha K_{rr}\sigma g_{rr}`$ $`=`$ $`0,`$ (28) $`{\displaystyle \frac{1}{2}}\beta ^r{\displaystyle \frac{dg_{\theta \theta }}{dr}}\alpha K_{\theta \theta }\sigma g_{\theta \theta }`$ $`=`$ $`0.`$ (29) The $`rr`$ and $`\theta \theta `$ components of equation (13) are respectively $`{\displaystyle \frac{d^2\alpha }{dr^2}}+{\displaystyle \frac{1}{2g_{rr}}}{\displaystyle \frac{dg_{rr}}{dr}}{\displaystyle \frac{d\alpha }{dr}}+\beta ^r{\displaystyle \frac{dK_{rr}}{dr}}+2K_{rr}{\displaystyle \frac{d\beta ^r}{dr}}`$ (30) $`+\alpha \left[{}_{}{}^{(3)}R_{rr}^{}+K_{rr}\left(2{\displaystyle \frac{K_{\theta \theta }}{g_{\theta \theta }}}{\displaystyle \frac{K_{rr}}{g_{rr}}}\right){\displaystyle \frac{\tau _{rr}}{2}}+{\displaystyle \frac{g_{rr}\tau _{\theta \theta }}{g_{\theta \theta }}}{\displaystyle \frac{\rho g_{rr}}{2}}\right]\sigma K_{rr}`$ $`=`$ $`0,`$ (31) $`{\displaystyle \frac{1}{2g_{rr}}}{\displaystyle \frac{dg_{\theta \theta }}{dr}}{\displaystyle \frac{d\alpha }{dr}}+\beta ^r{\displaystyle \frac{dK_{\theta \theta }}{dr}}+\alpha \left[{}_{}{}^{(3)}R_{\theta \theta }^{}+{\displaystyle \frac{K_{rr}K_{\theta \theta }}{g_{rr}}}+{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{\tau _{rr}}{g_{rr}}}\rho \right)g_{\theta \theta }\right]\sigma K_{\theta \theta }`$ $`=`$ $`0.`$ (32) The constraint equations $${}_{}{}^{(3)}R+K^2K_{ab}K^{ab}2\rho =0,D_aK_{}^{a}{}_{b}{}^{}D_bKJ_b=0,$$ (33) become in spherical symmetry $$\rho =\frac{{}_{}{}^{(3)}R_{rr}^{}}{2g_{rr}}+\frac{{}_{}{}^{(3)}R_{\theta \theta }^{}}{g_{\theta \theta }}+\frac{K_{\theta \theta }}{g_{\theta \theta }}\left(\frac{2K_{rr}}{g_{rr}}+\frac{K_{\theta \theta }}{g_{\theta \theta }}\right),$$ (34) $$2\frac{dK_{\theta \theta }}{dr}=g_{\theta \theta }J_r+\left(\frac{K_{rr}}{g_{rr}}+\frac{K_{\theta \theta }}{g_{\theta \theta }}\right)\frac{dg_{\theta \theta }}{dr}.$$ (35) ### B Generic coordinate conditions applied to spherical symmetry In spherical symmetry, the conformal strain lapse, equation (23) is given by $$\alpha =\frac{\frac{K_{rr}}{(g_{rr})^2}\left(\frac{1}{2}\beta ^r\frac{dg_{rr}}{dr}+g_{rr}\frac{d\beta ^r}{dr}\right)+\frac{K_{\theta \theta }}{(g_{\theta \theta })^2}\beta ^r\frac{dg_{\theta \theta }}{dr}\sigma \left(\frac{K_{rr}}{g_{rr}}+2\frac{K_{\theta \theta }}{g_{\theta \theta }}\right)}{\left(\frac{K_{rr}}{g_{rr}}\right)^2+2\left(\frac{K_{\theta \theta }}{g_{\theta \theta }}\right)^2}.$$ (36) The exponential curvature lapse, equation (25), is given by $`{\displaystyle \frac{d^2\alpha }{dr^2}}+\left({\displaystyle \frac{1}{2g_{rr}}}{\displaystyle \frac{dg_{rr}}{dr}}{\displaystyle \frac{1}{g_{\theta \theta }}}{\displaystyle \frac{dg_{\theta \theta }}{dr}}\right){\displaystyle \frac{d\alpha }{dr}}`$ (37) $`+\alpha g_{rr}\left[\left({\displaystyle \frac{K_{rr}}{g_{rr}}}\right)^2+2\left({\displaystyle \frac{K_{\theta \theta }}{g_{\theta \theta }}}\right)^2+{\displaystyle \frac{1}{2}}\left(\rho +{\displaystyle \frac{\tau _{rr}}{g_{rr}}}+2{\displaystyle \frac{\tau _{\theta \theta }}{g_{\theta \theta }}}\right)\right]`$ (38) $`+g_{rr}\left[\beta ^r{\displaystyle \frac{d}{dr}}\left({\displaystyle \frac{K_{rr}}{g_{rr}}}+2{\displaystyle \frac{K_{\theta \theta }}{g_{\theta \theta }}}\right)+\sigma \left({\displaystyle \frac{K_{rr}}{g_{rr}}}+2{\displaystyle \frac{K_{\theta \theta }}{g_{\theta \theta }}}\right)\right]=0.`$ (39) The minimal strain shift, equation (17), is given by $`{\displaystyle \frac{d^2\beta ^r}{dr^2}}+{\displaystyle \frac{d\beta ^r}{dr}}\left({\displaystyle \frac{1}{2g_{rr}}}{\displaystyle \frac{dg_{rr}}{dr}}+{\displaystyle \frac{1}{g_{\theta \theta }}}{\displaystyle \frac{dg_{\theta \theta }}{dr}}\right)`$ (40) $`+{\displaystyle \frac{\beta ^r}{2}}\left[{\displaystyle \frac{1}{g_{rr}}}{\displaystyle \frac{d^2g_{rr}}{dr^2}}+\left({\displaystyle \frac{1}{g_{\theta \theta }}}{\displaystyle \frac{dg_{\theta \theta }}{dr}}{\displaystyle \frac{1}{g_{rr}}}{\displaystyle \frac{dg_{rr}}{dr}}\right){\displaystyle \frac{1}{g_{rr}}}{\displaystyle \frac{dg_{rr}}{dr}}\left({\displaystyle \frac{1}{g_{\theta \theta }}}{\displaystyle \frac{dg_{\theta \theta }}{dr}}\right)^2\right]`$ (41) $`\left[{\displaystyle \frac{d}{dr}}\left({\displaystyle \frac{\alpha K_{rr}}{g_{rr}}}\right)+{\displaystyle \frac{\alpha }{g_{\theta \theta }}}{\displaystyle \frac{dg_{\theta \theta }}{dr}}\left({\displaystyle \frac{K_{rr}}{g_{rr}}}{\displaystyle \frac{K_{\theta \theta }}{g_{\theta \theta }}}\right)\right]=0.`$ (42) Thus, in spherical symmetry the generalized SY gauge is the combination of equations (37) and (40), while the generalized BCT gauge is the combination of equations (36) and (40). ### C Coordinate conditions tailored to spherical symmetry In spherical symmetry, one can also find simpler equations for the lapse and the shift. We note that equation (29) contains no derivatives of the shift, while equation (32) contains only one derivative of the lapse. We can use equation (29) to express the shift as $$\beta ^r=2\frac{\sigma g_{\theta \theta }+\alpha K_{\theta \theta }}{dg_{\theta \theta }/dr}.$$ (43) For $`\sigma =0`$ this is called the area freezing shift, as it keeps the area of a surface for constant $`r`$ constant. Similarly, we call (32) the area lapse. Substituting (43) into (32), we obtain a first-order differential equation for $`\alpha `$ alone. We can use the momentum and Hamiltonian constraints (35, 34) to simplify it to the form $$\frac{d\alpha }{dr}=\frac{g_{\theta \theta }}{dg_{\theta \theta }/dr}\left[\alpha \left({}_{}{}^{(3)}R_{rr}^{}\frac{2g_{rr}K_{\theta \theta }J_r}{dg_{\theta \theta }/dr}+\tau _{rr}+\rho g_{rr}\right)+2\sigma \left(K_{rr}\frac{g_{rr}g_{\theta \theta }J_r}{dg_{\theta \theta }/dr}\right)\right].$$ (44) We can integrate this first-order ordinary differential equation from the center of spherical symmetry outwards. However, the equation requires as a boundary condition the value of the lapse at one value of $`r`$. We refer to the combination of equations (43) and (44) as the area gauge. ### D Boundary conditions for the lapse and shift equations in spherical symmetry In three space-dimensions without spherical symmetry, some of the equations for the lapse and shift we have discussed are coupled elliptic equations. These require boundary values for the lapse and shift on a surface (topologically a 2-sphere) at large radius. One would expect the equivalent problem in spherical symmetry to be a second-order ODE boundary value problem, with one boundary condition each at $`r=0`$ and large $`r`$, for each of $`\alpha `$ and $`\beta ^r`$. One boundary condition at $`r=0`$ is immediately obtained from symmetry, namely that $`\alpha `$ must have an expansion in even powers of $`r`$ and $`\beta ^r`$ an expansion in odd powers of $`r`$. For the second boundary condition, it is as natural to specify $`\alpha `$ and $`lim(\beta ^r/r)`$ at $`r=0`$ as it would be to specify $`\alpha `$ and $`\beta ^r`$ at some large value of $`r`$. Doing that, we have the great simplification of integrating an ODE system from $`r=0`$ outwards, instead of solving a boundary value problem. The equations for $`\alpha `$ and $`\beta ^r`$ are homogeneously linear in $`\alpha `$, $`\beta ^r`$ and $`\sigma `$. The boundary values we need can be thought of as an overall factor in both $`\alpha `$ and $`\beta ^r/r`$, and their ratio. The ratio is uniquely determined at the center of spherical symmetry. In general we have $$\alpha K+D_a\beta ^a=\frac{(\mathrm{ln}\sqrt{g})}{t}=3\sigma ,$$ (45) where the first equality follows from the definition of $`K`$, while the second equality follows from (12). At the center of spherical symmetry, this yields $$\beta ^r=\left(\sigma +\frac{1}{3}K\alpha \right)r+O(r^3).$$ (46) The same result is also obtained by evaluating either (21), (23) or (43) at $`r=0`$. It remains to determine the overall factor in both $`\alpha `$ and $`\beta ^r`$. From (4) we have that $`\alpha =\alpha _0=\mathrm{const}.`$ at the origin in the Killing case. The value of $`\alpha _0`$ is a matter of convention that only affects the norm of the Killing vector field, and we can set it to any value that is constant in $`t`$. In the self-similar case, we formally have $`\alpha =\alpha _0\mathrm{exp}\sigma t`$ at the center, but this result is not very useful, as only one choice of $`\alpha _0`$ gives rises to a compatible coordinate system. To see this, label the central geodesic by proper time $`T`$ with the origin chosen so that $`T=0`$ is the singular point of the self-similar spacetime. In the self-similar time coordinate $`t`$ this point is $`t=\mathrm{}`$. Let $`T=T_0`$ where the initial slice $`t=0`$ intersects the central geodesic. Then it is easy to see that the proper relation $`T=T_0\mathrm{exp}\sigma t`$ is obtained only if we choose $`\alpha _0=\sigma T_0`$. In other words, we must guess correctly how far the slice is away from the singular point. A prescription for $`\alpha `$ at the center can be obtained by noting that we have $`_XT_{ab}=0`$ and therefore $`_X\rho =2\sigma \rho `$. In the 3+1 split this is $$(\alpha n^a+\beta ^a)_a\rho =2\sigma \rho .$$ (47) We combine this with the equation for $`n^a_a\rho `$ obtained from stress-energy conservation, $$0=n_a_bG^{ab}=K\rho +n^a_a\rho +D_aJ^a\tau ^{ab}K_{ab}+2J^aD_a\mathrm{ln}\alpha .$$ (48) At the center of spherical symmetry, $`D_a\alpha `$ and $`\beta ^a`$ both vanish, and we obtain an algebraic equation for $`\alpha `$ alone. Furthermore at the center we have $`K_{ab}=(K/3)g_{ab}`$ and a similar expression for $`\tau _{ab}`$, which simplifies our final result to $$\alpha =\frac{2\sigma \rho g_{rr}}{3\left[\frac{dJ_r}{dr}K_{rr}\left(\rho +\frac{\tau _{rr}}{g_{rr}}\right)\right]}+O(r^2)$$ (49) This equation gives a (nonzero) value for $`\alpha `$ at the origin provided that neither the numerator nor the denominator vanishes there. ## IV Tests in spherical symmetry ### A Numerical method We now discuss empirical tests of some of the coordinate conditions we have discussed for some given spacetimes, restricting ourselves to spherically symmetric examples. Rather than to evolve Cauchy data via the ADM equations, we evolve an embedded slice in the spacetime. All indices $`A,B,C,i,j,k`$ in this section denote coordinate components of tensors, rather than abstract tensor indices. Let a spacetime metric be given in closed form in coordinates $`x^A`$, giving rise to a metric $`g_{AB}`$. A three-dimensional spacelike slice embedded in the spacetime, together with a choice of three coordinates $`x^i`$ on the slice, is parameterized by four functions of three coordinates, $`x^A=F^A(x^i)`$. The four-metric $`g_{AB}`$ induces a three metric $`g_{ij}`$ and extrinsic curvature $`K_{ij}`$, and there is a unit timelike normal $`n^A`$ to the slice at each point. In our code, we choose a spacetime, and an initial slice, and calculate $`g_{ij}`$ and $`K_{ij}`$. The equations for doing this are derived in full in . If the spacetime contains matter, the components $`\rho `$, $`J_i`$ and $`S_{ij}`$ of the stress-energy tensor are also computed from $`T_{AB}`$. These are handed to a coordinate condition solver that returns $`\alpha `$ and $`\beta ^i`$. The slice is then evolved via $$\frac{F^A}{t}=\alpha n^A+\beta ^i\frac{F^A}{x^i}.$$ (50) We finite-difference the equations using a type of iterative Crank-Nicholson (ICN) algorithm. $`F^A/t`$ is computed at time $`t_n`$, and the $`F^A`$ are evolved to $`t_{n+1}`$ with a forward in time, centered in space (predictor) step. $`F^A/t`$ is then computed at time $`t_{n+1}`$, and the average of $`F^A/t`$ at $`t_n`$ and $`t_{n+1}`$ is then used for a corrector step. The corrector step is iterated to convergence. To maintain stability, we use numerical viscosity. In spherical symmetry, there are only two functions $`X^0(r)`$ and $`X^1(r)`$, and two components each of the induced metric and extrinsic curvature, $`g_{rr}(r)`$, $`g_{\theta \theta }(r)`$, $`K_{rr}(r)`$, and $`K_{\theta \theta }(r)`$, and stress-energy components $`\rho (r)`$, $`J_r(r)`$ and $`\tau _{rr}(r)`$. There is only one shift component, $`\beta ^r(r)`$, and the lapse, $`\alpha (r)`$. The coordinate condition solver and slice evolution code are independent and communicate only through the fields intrinsic to the slice, so that the coordinate condition solver never sees the $`F^A`$ directly. The code assumes that surfaces of constant $`t`$ are spacelike, and complains if, through a bad choice of coordinate, they are not, but no assumptions are made on the coordinates $`X^0`$ and $`X^1`$. They could, for example, be null coordinates. Integrating the embedding of a slice, rather than Cauchy data, has two advantages. The first one is stability: empirically, the ICN algorithm, together with a linear extrapolation outer boundary condition, is very stable for given $`\alpha `$ and $`\beta ^r`$. In particular, if we want to excise the center of spherical symmetry from the numerical domain in a black hole spacetime, an extrapolation inner boundary works well in the slice evolution code, while black hole excision in a genuine spacetime evolution code is a nontrivial problem even in spherical symmetry. The second advantage of the slice evolution is that we can use a given spacetime that is not a solution of the Einstein equations for any reasonable matter, but that has some properties worth investigating. In such a spacetime, we create a fictitious stress-energy tensor simply defined by the Einstein equations but not derived from any reasonable matter source. We think that such “fake” spacetimes can be valuable tests of qualitative aspects of coordinate conditions. ### B Stationary spacetimes We have used a variety of spacetimes to test three different gauges: area gauge, generalized SY gauge and generalized BCT gauge. The results are as follows. As model stationary spacetimes with a regular center we have examined Minkowski space, de Sitter space and the Schwarzschild solution for a constant density star, parameterized by $`R_{\mathrm{star}}/M_{\mathrm{star}}<9/4`$. All three gauge choices work on the Minkowski and de Sitter spacetimes. The area and generalized SY gauges work on the constant density star. The generalized BCT gauge fails numerically at the boundary of the star, where $`dg_{rr}/dr`$ is discontinuous. As a boundary condition we choose $`\alpha =1`$ at the center. As the equations are linear in both $`\alpha `$ and $`\beta ^r`$, this choice is equivalent to any other choice that is constant in time. As a consistency check of our programming, we have verified that the Hamiltonian and momentum constraints are obeyed up to finite differencing error. As the initial slice, we take $`X^1=r`$ and $`X^0(r)`$ a Gaussian, with $`X^0`$ and $`X^1`$ the standard time and radial coordinates. We find that the induced metric, extrinsic curvature, stress-energy components, lapse and shift are all time-independent up to numerical error. To test coordinate conditions in a black hole spacetime with the black hole excised from the numerical domain, we have implemented the Schwarzschild spacetime in coordinates that are regular at the future event horizon (the Painlevé-Güllstrand coordinates). Here the area gauge and generalized SY gauge work. The generalized BCT gauge has numerical problems at the excision boundary. As a boundary condition we have chosen $`\alpha =1`$ at the excision boundary, which is chosen at constant coordinate $`r`$. As the shift is supposed to keep our generic radial coordinate $`r`$ at constant curvature radius $`\sqrt{g_{\theta \theta }}`$, this boundary condition is consistent. As the initial slice, we took $`X^1=r`$ and $`X^0(r)`$ a Gaussian, with $`X^1`$ the area radius and $`X^0`$ the Painlevé-Güllstrand time (which is defined by surfaces of constant $`X^0`$ being flat and spacelike). Again we find that all variables are constant in time, even though they are all nontrivial. ### C Self-similar spacetimes We have also investigated four self-similar spacetimes. Our coordinate prescriptions are designed to bring the metric of a CSS spacetime into the form $$g_{\mu \nu }=e^{2t}\overline{g}_{\mu \nu }$$ (51) where the components of $`\overline{g}_{\mu \nu }`$ are independent of $`t`$. The metric of a DSS spacetime can be written in the same form, with $`\overline{g}_{\mu \nu }`$ periodic in $`t`$. It is important to note that while all our prescriptions must work on CSS spacetimes, barring numerical error and numerical instability, it is not guaranteed that they work, in the sense of bringing the metric into this form, on DSS spacetimes. The reason is that given a CSS spacetime and an initial slice, the coordinate system of this form is unique, while it is not for a DSS spacetime. In order to see how successful a gauge is in making the conformal spacetime metric independent of $`t`$ or periodice in $`t`$, it is useful to rescale fields intrinsic to the slice ($`g_{ij}`$, $`K_{ij}`$ etc.) with appropriate powers of $`\mathrm{exp}t`$ so that the rescaled fields are independent of time if and only if the spacetime is CSS and our algorithm finds the conformal Killing vector. The simplest CSS spacetime, Minkowski spacetime, is not suitable for our investigation, as the homothetic vector is not unique. Instead, we have considered a conformally flat spacetime of the form $$ds^2=e^{2t}\left(dt^2+dr^2+r^2d\mathrm{\Omega }^2\right).$$ (52) We have given this spacetime fictitious matter defined by the Einstein equations. It has a unique homothetic vector field if by assumption we restrict that vector field to the $`tr`$ plane. Here, we have tested all three gauges. All work well when we use the the analytic boundary condition $`\alpha =e^t`$ at the center. A less trivial example is the Roberts solution, a family of CSS massless scalar field solutions, which in double null coordinates is given by $`ds^2`$ $`=`$ $`dudv+{\displaystyle \frac{1}{4}}\left[(1p^2)v^22uv+u^2\right]d\mathrm{\Omega }^2`$ (53) $`\varphi (u,v)`$ $`=`$ $`\mathrm{ln}{\displaystyle \frac{(1p)vu}{(1+p)vu}},`$ (54) with $`p`$ a parameter. This spacetime has curvature singularities at $`g_{\theta \theta }=0`$. We have investigated the case $`p=0.5`$, where these singularities are naked, and lie in the future and past lightcones. The slice we evolve lies in the spacelike sector. The only boundary condition for $`\alpha `$ at the inner boundary that worked here is $`\alpha =\alpha _0\mathrm{exp}\sigma t`$, with $`\alpha _0`$ taken from the exact solution. Here the area gauge and generalized SY gauge work. The generalized BCT gauge has numerical problems at the excision boundary. Our other two examples are of more physical interest. They are the CSS critical solution for the collapse of a perfect fluid with the equation of state $`p=(1/3)\rho `$ and the DSS critical solution for the collapse of a massless scalar field . These solutions have been constructed numerically as solutions of a nonlinear ODE eigenvalue problem , and a nonlinear PDE eigenvalue problem , respectively. We interpolate them numerically to the required values of $`X^0`$ and $`X^1`$. For both spacetimes, we have tested the area gauge and the generalized SY gauge. Again, we use a Gaussian initial slice. In the CSS fluid spacetime, we have been able to use the boundary condition (49). The boundary condition works, and all variables are constant in time. The DSS scalar field spacetime has a regular center, but the scalar field is a periodic function of $`\tau `$ at the center, so that the density $`\rho `$ vanishes twice per period. This means that we cannot apply the boundary condition (49). Instead, we have applied the exact boundary condition $`\alpha =\alpha _0\mathrm{exp}\sigma t`$, with $`\alpha _0`$ taken from the exact solution. This test is nontrivial, as it is not clear at all that a coordinate condition designed to catch a continuous symmetry can catch a discrete one. But it works: all variables are periodic in time for the scalar field DSS spacetime for at least two periods, in both the area and generalized SY gauge. ## V Conclusions We have examined the problem of finding good coordinates in numerical relativity, and the problem of defining a renormalisation group for general relativity in critical collapse, in a single framework. We have listed old and new prescriptions that obey our criterion for a good coordinate condition in either application: that the presence of an approximate timelike Killing field, or an approximate homothetic vector field, in some region of the spacetime (for example at late times) be revealed by metric coefficients that vary only slowly with time. We have tested several prescriptions numerically in spherically symmetric spacetimes that are exactly stationary or homothetic. For these exact symmetries, it is clear analytically that our prescriptions will work, but we have used these cases both to test our code, and to see how well different schemes can be implemented numerically. We find that a scheme specific to spherical symmetry (area freezing shift and radial lapse) is the most stable one, generalized Smarr-York (SY) gauge (minimal strain shift and exponential $`K`$ lapse) comes second, while we have been unable to implement the generalized Brady-Crighton-Thorne (BCT) gauge (minimal strain shift and conformal strain lapse) stably. What this means is that the combination of our numerical methods of solving for the BCT gauge and of evolving a slice is numerically unstable. This indicates that coupling the BCT gauge to a genuine spacetime evolution code may also be numerically unstable, but it is of course possible that a different numerical implementation of the gauge and/or the evolution equations is stable. Other difficulties of the BCT gauge are evident analytically: Not all initial data sets (in our case, initial slices through a given spacetime) admit a solution to BCT gauge, and BCT gauge may still break down (when $`K_{ab}K^{ab}=0`$ anywhere) at a later time. In spherical symmetry, the slice must nowhere be close to a polar slice for BCT gauge to work. We have entered new terrain examining one spacetime (the Choptuik scalar field critical solution) with a discrete, rather than continuous, symmetry. Here we found numerically that two prescriptions made for a continuous symmetry (area gauge and generalized SY gauge) actually follow the discrete symmetry, in the sense that the metric becomes periodic in time. (Our third gauge, generalized BCT gauge, was unstable numerically in our implementation, but may yet work with better numerical methods.) This is an important step in the construction of a renormalisation group for critical collapse, as it suggests that a single coordinate prescription may realize the renormalisation group both for continuously and discretely self-similar (CSS and DSS) critical solutions. It is also encouraging for the use of co-rotating coordinates in non-circular binary black hole inspiral. One crucial ingredient is still lacking in a working prescription for a renormalisation group, namely the boundary conditions used when solving elliptic equations for the lapse and shift. Details aside, these boundary conditions must contain a good guess for how far away in proper time the slice is from the accumulation point of the (approximately) self-similar spacetime. Although there are ideas \[see equation (49)\], we have not been able to solve this problem, and instead had to supply this information by hand. Future work should include testing beyond spherical symmetry and in particular with angular momentum, and on spacetimes with approximate symmetries, as well as developing ideas on what boundary conditions should be imposed on elliptic coordinate conditions. ## ACKNOWLEDGMENTS We would like to thank Beverly Berger, Matt Choptuik, Bob Wald and Stu Shapiro for helpful discussions. DG would like to thank the Albert Einstein Institute and the University of Chicago for hospitality. DG and CG would like to thank the Institute for Theoretical Physics at Santa Barbara (partially supported by NSF grant PHY-9407194) for hospitality. This work was partially supported by NSF grant PHY-9514726 to the University of Chicago and by NSF grant PHY-9722039 to Oakland University.
no-problem/9908/astro-ph9908315.html
ar5iv
text
# Simulations of Nonthermal Electron Transport in Multidimensional Flows: Application to Radio Galaxies ## 1 Introduction The jet-based, dynamical paradigm for radio galaxies now seems secure, but our understanding of the basic physics remains primitive. A key barrier has been our limited ability to fill the enormous gap between complex plasma flows thought to drive the phenomenon and the emissions that reveal them. The emissions come from particles very far from thermodynamic equilibrium, so that their energy distributions depend on local microphysics and their histories. Further, the posited plasma flows are strongly driven systems, so they are also mostly far from any stable dynamical equilibria. Among other things this leads to very unsteady flows. These characteristics make numerical simulations a necessity to push beyond the simplest characterizations of the phenomena and to verify even simple analytical calculations based on equilibrium assumptions. The past decade has seen a revolution in our ability to carry out sophisticated multidimensional gasdynamical simulations, including magnetic fields (e.g., Clarke et al. 1989; Lind et al. 1989; Kössl et al. 1990; Nishikawa et al. 1998). The diffusive shock acceleration paradigm for particle energization, which is relatively robust in relating bulk dynamics and particle spectra, seems to offer an attractive way past ignorance of some microphysical details. Despite these advances, modeling the relevant particle microphysics adequately in simulations has remained illusive. These difficulties come from both severe technical constraints on simulations and ignorance of some of the most important physical parameters. Now we have initiated a program that takes a significant technical step forward, and which we hope will help resolve some of the key physical ambiguities. In multi-dimensional MHD flow simulations we follow the nonthermal electron population in some detail, allowing us to compute for the first time meaningful model emissions. This paper describes some initial results as they apply to the properties of the nonthermal electrons. A companion paper (Tregillis et al. 1999) describes application of our methods to synthetic radio and X-ray observations and what they seem to be telling us about interpretation of observations in terms of physical source parameters. A report of initial 2-D simulations, plus additional references are given in Jones et al. 1999. The 3-D results shown here are preliminary. A full report is in preparation. ## 2 Background Our flow dynamics is treated with a second-order accurate, conservative, “TVD” ideal MHD code described in Ryu & Jones 1995; Ryu et al. 1995; Ryu et al. 1998. It maintains the divergence free condition to the magnetic field to machine accuracy using an upwinded constrained transport scheme as described in Ryu et al. 1998. For energetic particle transport we use the conventional “convection diffusion equation” for the momentum distribution function, $`f`$, (e.g., Skilling 1975) which follows spatial and momentum diffusion as well as spatial and momentum advection of the particles. The last of these corresponds to energy losses and gains from, for example, adiabatic expansion and synchrotron aging. However, high computational costs prohibit solving this equation through standard finite difference methods in complex flows expected in radio galaxies. To circumvent this we use a conservative finite volume approach in the momentum coordinate, taking advantage of the broad spectral character expected for $`f(p)`$. Particle fluxes across momentum bin boundaries are estimated by representing $`f(p)`$ as $`f(p)p^{q(p)}`$, where $`q(p)`$ varies in a regular way. Numerically we use the integrated number of electrons within each bin and the slope, $`q`$, within each bin. Thus, we can follow electron spectral evolution in smooth flows for all the effects mentioned above with a modest number of momentum bins. Typically we have used 8 bins to cover energies up to a few hundred GeV for electrons. We have shown that this approach produces solutions in good agreement with more conventional methods, including diffusive acceleration at shocks and synchrotron energy losses. However, in the flows being studied, diffusive acceleration of electrons to GeV energies at shocks is effectively instantaneous within a dynamical time step. Because of that direct simulation of this physics would be prohibitively expensive. We can, however, also circumvent this difficulty if we assume the analytic, steady, test particle form for the electron distribution just behind shocks. Ignoring for this discussion some details, that spectrum will be a power law with an index, $`q=\frac{3r}{r1}`$, where $`r`$ is the shock compression ratio. Together these features give us a powerful tool for numerical simulations of such complex phenomena as radio galaxies (Jones et al. 1999; Jun & Jones 1999). Our method is complementary to that recently described by Micono et al. 1999. Those authors followed the convection-diffusion equation in detail outside of shocks using conventional methods, but were constrained by the numerical limitations we describe to follow a very limited number of Lagrangian volume elements within the flow. That approach does not allow one to make synthetic observations of the simulated flows, whereas ours does. ## 3 Discussion Our initial radio galaxy explorations with this new tool have focussed on improved understanding of the ways that complex jet flow dynamics influence the spectral properties of electrons. For this purpose we carried out 2-D axisymmetric and fully 3-D MHD simulations of light, pressure-balanced, supersonic jets carrying weak, but “active”, magnetic fields, which are helical inside the jet and axial in the uniform ambient medium. So far we have followed the dynamics until the jets propagate $`4050`$ jet radii, considering cases where the magnetic fields are strong enough to cause significant synchrotron aging and cases where it is too weak for this to matter. We have considered cases in which the relativistic electrons are injected naturally as part of diffusive acceleration physics at shocks and cases where only an electron population introduced with the jet flow is represented. For the latter cases, we assume this population comes onto the grid with a power law spectrum, $`q=4.4`$, corresponding to a synchrotron spectral index, $`\alpha =\frac{q3}{2}=0.7`$. In all cases the spectra evolve appropriately within the flows in response to shock acceleration, synchrotron aging and adiabatic effects. Fig. 1 illustrates some of the important common properties in our simulations. This case is a 3-D, Mach 8 (internal Mach number), light, MHD jet ($`\rho _j/\rho _a=10^2`$, $`\beta =p_b/p_g10^2`$, where $`\beta `$ refers to the magnetic field on the jet axis.), which was slowly precessed on a 5 degree cone at the origin. The center of the jet origin is marked in the figure by the small green boxes. This numerical experiment was carried out on a $`576\times 192\times 192`$ grid. The inflowing jet had a top hat velocity profile with a thin transition sheath around it. The core radius spanned 15 zones. This flow is not at all steady, since the jet terminus tends to “whip” around, sometimes forming “splatter spots” (Williams & Gull 1985) and then pinching off and redirecting itself. Here we show properties at a moment when the flow is relatively simple. The left panel reveals through volume rendering of $`𝐮`$ the locations of shocks in the jet and its back-flow. The strongest shocks appear blue, while relatively weaker shocks are red. The right panel shows the spatial distribution of the spectral index of $`10`$ GeV electrons at the same time. Synchrotron aging is negligible in this case, so we are seeing in the electron spectra only effects from shock acceleration and advective mixing. Shock-modified spectra flatter than the “injected spectrum” are shown in color, with the relatively flatter (steeper) spectra being blue (red). The injected spectral index, $`q=4.4`$, is rendered in white. Given the absence of synchrotron aging, all spectra are at least as flat as that entering at the jet origin. For this simulation there were no additional electrons injected at shocks; that is, the entire relativistic electron population entered with the jet. In this case those electrons constitute a fraction $`10^4`$ of the total number of electrons. The remainder are thermal. Since the jet speed is assumed to be 0.1 c, thermal electrons in the jet have energies $`70`$ keV. Only flow regions filled with plasma originating in the jet are shown in the figure. So, for example, the bow shock preceding the jet is not visible in the left panel, since it occurs in the ambient medium. These images illustrate dramatically that shock structures in jet driven flows are very complex, and also show our consequent key finding that the spectral properties of electrons emerging into the cocoon are extremely heterogeneous. Comparison of the images shows us why this finding makes sense. First, let us locate the jet flow itself. Near its base the jet can be followed through its conical internal shocks. The terminus of the jet occurs at the far upper left in a small, strong shock. It turns out that only the central core of the jet actually exits through that terminal shock. Much of the plasma emerging from the jet has passed through only weak shocks before it is redirected into the cocoon. Consequently, shocks inside and at the end of the jet have had a relatively small influence on the spectra of electrons entering the cocoon. Note next that most of the rather complex “shock web” near the jet head involves cocoon flow, in fact. Remarkably, the strongest shocks are often in the cocoon, rather than the jet. They appear to be generated by the non-axial motions of the jet head. The distribution of electron spectra in the adjacent image does not map in an obvious way onto the shock distribution. Some insight into this complication comes from noting the apparent “streams” in the particle spectra, which are especially apparent near the jet terminus. The streams highlight flows downstream of localized, strong shocks. This feature emphasizes that the shocks are themselves very complex, and also that we are seeing a blend of many different flow histories. The relative importance of shocks inside the cocoon compared to the jet can be recognized by understanding the origin of a relatively flat spectral region visible in the figure roughly 2.5 cm to 4.5 cm from the end of the jet towards the origin (green box). There, $`q4.3`$, so it shows yellow in the image. Examination of flow streaklines shows that this plasma all passed through a small shock visible in the shock image (with a yellow color) about 2.5 cm from the end of the jet, but physically in the cocoon, not the jet itself. These results emphasizing the complexity of the evolution of particle spectra in jet head regions support and augment our earlier findings from 2-D axisymmetric simulations (Jones et al. 1999). An analogous finding from the 2-D study was that magnetic field structures are highly inhomogeneous (“intermittent”), especially in the head region. Thus, synchrotron emissivity and associated aging rates are very inhomogeneous. That behavior seems even stronger in our initial examination of a 3-D run that was similar to the one shown in Fig. 1, except that the magnetic field was strong enough to produce significant aging on the time scales simulated. Observational properties of these flows are described in the companion paper by Tregillis et al. 1999. This work was supported at the University of Minnesota by the NSF through grant AST96-16964 and by the University of Minnesota Supercomputing Institute. DR was supported in part by KOSEF through grant 975-0200-06-2. Q: D. Harris Great code, but deductions about particle vs B field distributions may be misleading, because you still have MHD, where basic fluid controls where and how much e & B occur/get carried. Is it not true that in real jets the dynamics may be controlled by the relativistic particles and B field? A: T. Jones Magnetic fields are fully dynamically coupled in our simulations. In fact, although we described the magnetic fields as “weak” because the nominal ratio of gas to magnetic pressure in the jet is large, we find that there are numerous places in the cocoon where magnetic, Maxwell stresses are significant. They are properly accounted for. The relativistic particles represent a high energy tail to the bulk population, so their dynamics accurately reflects the total. We have restricted their energy content so that it does not dominate in order that the overall flows may be treated non-relativistically. The relativistic particles are handled correctly in the context of the nonrelativistic bulk flow. Eventually we will want to do simulations like these with fully relativistic MHD codes. That step is some distance off, even if we accept the good progress being made on relativistic fluid codes. Treatment of particle acceleration at relativistic shocks is much more subtle than at nonrelativistic shocks, so it is not obvious how we will make that step for complex problems. Here we have assumed that the shocks are not relativistic and not superluminal.
no-problem/9908/nucl-ex9908003.html
ar5iv
text
# Neutron Electric Form Factor from ³("He")⃗⁢(𝑒⃗,𝑒'⁢𝑛) ## 1 Introduction The electric form factor of the neutron has been relatively poorly known for a long time, which is mainly due to the absence of free neutron targets, so one has to resort to light nuclei as effective neutron targets. This complicates the interpretation of the experiment, since nuclear structure and proton contributions need to be taken into account. E.g. the Platchkov inclusive scattering data on Deuterium have an impressive statistical accuracy, but the $`G_E^n`$ extraction depends strongly on the choice of the $`NN`$-potential used for the deuteron wave function: the absolute magnitude varies over 50%. Double polarized experiments have been proposed because this model dependency and experimental systematic uncertainties cancel out to a great extent by measuring asymmetries. Moreover, in exclusive experiments the results depend less on the proton contributions; but on the other hand become more sensitive to final state interactions. The general expression for the cross section of reactions with polarised beam and target is : $$\sigma =\sigma _0\times \left\{1+𝐀^\mathrm{𝟎}𝐒+h\left(A_e+𝐀^{}𝐒\right)\right\}$$ where $`𝐒`$ and h are the target polarisation and the electron helicity; $`𝐀^\mathrm{𝟎}`$ and $`A_e`$ are the target and electron analyzing powers and $`𝐀^{}`$ the spin-spin correlation functions. The Cartesian coordinate frame for the vectorial polarisation observables has x̂ in the scattering plane, perpendicular to the 3-momentum transfer $`𝐪`$, ŷ perpendicular to the scattering plane and ẑ parallel to $`𝐪`$. On a free neutron (and in a nucleus in PWIA) $`A_x^{}`$ for $`(\stackrel{}{e},e^{}n)`$ is proportional to $`G_E^n/G_M^n`$ and $`A_z^{}`$ in the inclusive channel is proportional to $`G_M^n`$, while all other components of $`𝐀^\mathrm{𝟎}`$ and $`𝐀^{}`$ are zero. In our experiment we chose $`{}_{}{}^{3}\mathrm{He}`$ as an effective neutron target. Since for 90% of the $`{}_{}{}^{3}\mathrm{He}`$ ground state wave function the two protons are in a relative $`S`$ state, $`{}_{}{}^{3}\stackrel{}{\mathrm{He}}`$ may be considered as an effective polarised neutron target. However, the $`D`$ state and the $`S^{}`$ state still contribute about 8% and 2%, respectively, which can not be neglected since their contribution to $`A_x^{}`$ is as important as that of $`G_E^n`$. Another complicating factor is the influence of the final state interactions. Their importance was demonstrated by the measurement of $`A_y^0`$ in the first run of the 9405 experiment in 1997, where this observable, which is zero in PWIA, was found to be $`0.50\pm 0.05`$ at $`Q^2=0.2\text{ GeV}^2/c^2`$ . This suggests that an accurate theoretical description, that takes all relevant reaction reaction mechanisms consistently into account, is needed for a reliable interpretation of the data, in particular for the extraction of $`G_E^n`$. We compare our data for $`(e,e^{}p)`$, $`(e,e^{}d)`$ and $`(e,e^{}n)`$ with calculations Laget , Nagorny and Golak (Glöckle *et al.* ). ## 2 Experimental setup ### 2.1 Polarised electrons The experiment was performed in the internal target hall of the MEA/AmPS facility at NIKHEF. Pulses of $`2.1\mu \text{s}`$ of polarised electrons were produced by means of photoemission from strained layer III-V semiconductor crystals in an ultrahigh vacuum electron gun . The spin of these 100 keV electrons could be directed (using electrostatic bends and solenoids) and measured (with a Mott polarimeter). The electrons were accelerated by MEA to 720 MeV and injected into the AmPS storage ring. A ’Siberian Snake’, consisting of a set of superconducting solenoids, preserved the electron polarisation during storage. The electron polarisation in the target area was measured with a Compton Backscattering Polarimeter . It was possible to store over 200 mA, but we observed that the beam polarisation was lost when more than 140 mA was injected. During the experiment a maximum injection current of 120 mA was used. ### 2.2 Polarised $`{}_{}{}^{3}\mathrm{𝐇𝐞}`$ In the target area, the beam is led through a Ø 15 mm, 40 cm long cylindrical storage cell. Polarised $`{}_{}{}^{3}\text{He}`$ gas is bled into the storage cell with a flow of $`10^{17}\mathrm{atoms}/s`$ resulting in a density of $`10^{15}\mathrm{atoms}/\mathrm{cm}^2`$ at a temperature of 30 K. The $`{}_{}{}^{3}\text{He}`$ gas was polarised in a pumping cell directly above the target with a metastability-exchange optical pumping technique. Transitions to the $`2^3S_1`$ state were induced with an RF generator, after which illumination with circularly polarised laser light of 1083 nm excited the atoms to a $`{}_{}{}^{2}P_{0,1,2}^{}`$ state, thereby depleting the $`2^3S_1`$ hyperfine levels with the nuclear spin (anti-)parallel to the target holding field. The polarisation is then transferred to the atoms in the ground state via metastability-exchange collisions. The nuclear polarisation was measured by monitoring the fluorescence light of the $`3^1D_22^1P_1`$ transition. ### 2.3 Detectors The scattered electrons were detected with the BigBite spectrometer , consisting of a 1 Tesla dipole magnet, two sets of drift chambers for tracking and a scintillator and Čerenkov detector for triggering and particle identification. The momentum bite is 250-900 MeV/c with a resolution of about $`0.5\%`$, while a solid angle extended over $`10^{}`$ in the polar angle (resolution $`0.2^{}`$, central angle $`40^{}`$) and $`34^{}`$ in the azimuthal angle (resolution $`0.2^{}`$). Protons and deuterons were detected with the Range Telescope, consisting of two wire chambers, a hodoscope and fifteen 1 cm thick scintillator layers. Protons with an energy of 35-150 MeV and deuterons with an energy of 45-200 MeV could be detected with a resolution of 1 MeV in energy and $`1^{}`$ in angle. Neutrons were detected in two consecutive time-of-flight walls, positioned at 2.2 m and 3.1 m (respectively) from the centre of the target at a central angle of $`56^{}`$ with the electron beam. Each wall consists of four horizontally stacked telescopes, each telescope consisting of a $`20\times 20\times 160\mathrm{c}\mathrm{m}^3`$ scintillator bar preceded by 1.0 cm and 0.3 cm thick veto layers. All layers have double-sided readout, enabling position determination with a resolution of about 5 cm. The time of flight of the neutrons from the interaction vertex to the neutron detector (ND) was deduced from the trigger time difference of BigBite and the ND, corrected for the electron time-of-flight and for the time of flight of the scintillator light from the impact position to the triggering photomultiplier. Detection of protons in elastic $`\mathrm{H}(e,e^{}p)`$ determined the offset in the time-of-flight. The resolution of 1.0 ns (sigma) was limited by the jitter in the triggering scintillator of BigBite. Neutral particles were selected by requiring no hit in the veto layers of the triggering scintillator bar and those of the neighbouring bars. A plot of deposited energy versus time for neutral particles is shown in figure 2. The band at 8 ns is associated with pion production. It disappears when $`\omega <170\text{ MeV}`$ is required. Since the random coincidence events have an almost flat distribution in $`E_m`$ over a range of severel hundred MeV, most of this background can be rejected by placing a cut on $`E_m`$. ## 3 Results In figure 2 $`A_x^{}`$ is plotted versus the missing momentum ($`p_m=|𝐪𝐩_n|`$), all other kinematical degrees of freedom are averaged over the acceptance, within the following ranges: $$\begin{array}{ccccc}0.14\text{ GeV}^2& <& Q^2& <& 0.26\text{ GeV}^2\\ 40\text{ MeV}& <& \omega & <& 170\text{ MeV}\\ 550\text{ MeV}& <& E^{}& <& 680\text{ MeV}\\ 10\text{ MeV}& <& E_m& <& 40\text{ MeV}\\ 200\text{ MeV}& <& p_n& <& 600\text{ MeV}\end{array}$$ with $`E_m=\omega T_nT_{pp}`$, where $`T_n`$ is the neutron kinetic energy and $`T_{pp}=\sqrt{p_m^2+4m_p^2}2m_p`$ is the kinetic energy of the center of mass of the recoiling two protons. The resolution in $`E_m`$ was 13 MeV. Figure 2 contains about 50-70% of the final statistics. The analysis is ongoing and this result is very preliminary. One-loop calculations of Nagorny were performed for a grid of 50000 points, which was used by a Monte Carlo program for interpolation. The resulting curves for $`G_E^n=0`$ and $`G_E^n`$ equal to the Galster fit are plotted seem to suggest that this data set will result in a value between zero and one times the Galster fit. The two-loop calculations differ from the one-loop calculations at low $`Q^2`$; this needs to be investigated further.
no-problem/9908/cond-mat9908388.html
ar5iv
text
# Role of confined phonons in thin film superconductivity \[ ## Abstract We calculate the critical temperature $`T_c`$ and the superconducting energy gaps $`\mathrm{\Delta }_n`$ of a thin film superconductor system, where $`\mathrm{\Delta }_n`$ is the superconducting energy gap of the $`n`$-th subband. Since the quantization of both the electron energy and phonon spectrum arises due to dimensional confinement in one direction, the effective electron-electron interaction mediated by the quantized confined phonons is different from that mediated by the bulk phonon, leading to the modification of $`T_c`$ in the thin film system. We investigate the dependence of $`T_c`$ and $`\mathrm{\Delta }_n`$ on the film thickness $`d`$ with this modified interaction. PACS Number : 63.22.+m, 74.20.-z, 74.25.Kc \] Superconductivity in thin films has been studied for the last four decades. The phenomenon of thin film superconductivity has its own specific peculiar features. In early investigations, the effect of the film size on the superconducting transition temperature, $`T_c`$, for thin films was investigated. Experiments have shown a monotonic increase of the critical temperature, $`T_c`$, with decreasing film thickness. From the theoretical point of view, the shape resonances and the strong thickness dependence of $`T_c`$ are the characteristic features of thin film superconductivity. The size quantization of the transverse motion of the electron in the film leads to an increase of $`T_c`$ with decreasing film thickness, arising essentially from an enhanced effective BCS pairing interaction. The resonance effects are manifest each time one of the $`n`$ two-dimensional (2D) subband energy levels, $`E_n(d)`$, of a film with thickness $`d`$, (for ‘transverse’ motion perpendicular to the film faces or along the confinement direction), passes through the Fermi surface as the thickness $`d`$ is varied. In previous calculations, the phonon modes were assumed to be the same as in the bulk material and only the one-dimensional quantum confinement effects of electrons were considered; i.e., the superconductivity in the thin film was considered to be arising from the attractive electron-electron interaction mediated by bulk phonons interacting with subband-quantized electrons. However, phonons in thin films (‘slab’ phonon modes) also have specific characteristic features by virtue of definite boundary conditions imposed by confinement in thin films. The phonon dispersion in a thin film undergoes substantial modification compared with the bulk, and a quantization of the phonon spectrum occurs . The quantization of the phonon spectra has many effects on thin film electronic properties, which have been extensively studied, particularly in the context of semiconductor quantum well structures. It is the purpose of this paper to reconsider the previous calculations of thin film superconductivity taking into account confined size-quantized phonons in the thin film. We find the resonant shape of the superconducting transition temperature, $`T_c`$, arises from both the quantum electronic confinement and phonon confinement. Following the BCS theory the critical temperature $`T_c1.14\omega _D\mathrm{exp}(1/N_0V_0)`$ for a bulk BCS superconductor, where $`\omega _D`$ is the phonon Debye frequency, $`N_0`$ the electronic density of states at the Fermi energy, and $`V_0`$ the effective attractive electron-electron interaction mediated by phonon exchange. An increase of $`N_0V_0`$ implies an increase in $`T_c`$. For a thin film with thickness $`d`$, the density of states at the Fermi energy develops quantum structure due to the confined 2D subbands: $`N_{film}=(2\nu 1)m/4\pi d=N_{3D}(2\nu +1)\pi /2dk_F`$, where $`\nu =1,2,\mathrm{}`$ is the occupied subband index. Thus, when we use the bulk electron-phonon coupling constant $`J`$ unmodified by any phonon size confinement corrections, which is identical for all subbands, the critical temperature for thin films depends on the film thickness, and decreases exponentially as the thickness increases. However, for a fixed bulk electron density (or, $`k_F`$), as the film thickness increases the higher quantized subbands are occupied by Cooper pairs, so that the critical temperature of the film jumps to a higher value due to the higher $`\nu `$, arising entirely from the jumps in the density of states as the effective Fermi level moves through higher values of the subband index $`\nu `$. This implies that with increasing thickness of the film the critical temperature of the film exhibits resonance features. In real thin films, therefore, the electron-phonon coupling constant is different from that in the bulk material due to the quantization of the phonon dispersion , and therefore the simple resonance scenario (discussed above) must be modified. When we consider the confined phonons in the thin film, the electron-phonon coupling constant is different for different subbands, and the effective coupling strength decreases with increasing subband index; that is, higher subbands have progressively weaker electron-phonon coupling constants. This change of the coupling constant gives rise to features in the superconducting energy gap of thin films, which have not earlier been considered in the literature. In this paper we calculate the critical temperature, $`T_c`$, and superconducting energy gaps, $`\mathrm{\Delta }_n`$, with the modified electron-electron interaction mediated by the confined slab phonons. The results for the calculated critical temperature obviously depend on the slab phonon dispersion, which in turn depends on the boundary conditions used in the phonon confinement model. We have followed here the simple approach of Thompson and Blatt , where the boundary has been treated by an infinite wall, and we use wavefunctions which vanish at the boundary. Although this is a highly simplified phonon confinement model it has the virtue of being analytically tractable — one could systematically improve upon this model using our theory as the starting point. We expect our results to be well-valid qualitatively. We assume that our superconducting film with a finite width, $`d`$, is confined in the $`z`$ direction by an infinite square-well confinement potential applied at $`z=0`$ and $`z=d`$. We choose the same infinite (one dimensional) square-well confinement for both electrons and phonons, as would be appropriate for a free-standing thin film. Using periodic boundary conditions in the $`x`$ and $`y`$ directions with periodicity distance $`L`$, we have the one-electron wave function and spectrum $`\varphi _{𝐤,n}(𝐫,z)`$ $`=`$ $`u_n(z)\mathrm{exp}(i𝐤𝐫)/L,`$ (1) $`ϵ_n(𝐤)`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2k^2}{2m}}+E_n.`$ (2) Here, $`𝐤=(k_x,k_y)`$, $`𝐫=(x,y)`$ are the 2D wave vector and position vector in the plane of free motion, and by solving Schrödinger equation for $`z`$-direction with confinement potential we have $`u_n(z)`$ $`=`$ $`(2/d)^{1/2}\mathrm{sin}(k_nz),`$ (3) $`E_n`$ $`=`$ $`{\displaystyle \frac{\mathrm{}^2(k_n)^2}{2m}},`$ (4) where $`k_n=n\pi /d`$ with $`n=`$ integer. Thus, confinement in the $`z`$ direction leads to the quantization of electron energy levels into different subbands. In addition to the quantization of electron energy, we take into account the modification in the thin film phonon dispersion arising from the quantization of the phonon spectrum . The quantization of the phonon spectra leads to the change of the conventional electron-phonon interaction. The specific expression for the electron-phonon coupling in the thin film can be obtained on the basis of the general deformation potential electron-phonon interaction theory $$H_{\mathrm{ep}}=Dd^3r\mathrm{\Psi }^{}(r)w(r)\mathrm{\Psi }(r),$$ (5) where $$\mathrm{\Psi }(r)=\frac{1}{\sqrt{A}}\underset{k,n}{}u_n(z)e^{ikr}c_{kn},$$ (6) $`A`$ is the 2D area, $`c_{k,n}`$ is the destruction operator for an electron with momentum $`k`$ in the n-th subband, $`w(𝐫)`$ is the lattice displacement vector, $`D`$ is the deformation potential coupling constant (which we uncritically assume to be the bulk value since not much experimental information is available). The displacement vector $`w(𝐫)`$ can be obtained by solving equations of motions of the elastic continuum with the appropriate boundary conditions . We skip the details of this elastic continuum theory (which involves substantial algebra) for the sake of brevity Within the framework of the BCS theory, Cooper pairs are produced consisting of electrons with opposite k and identical n in a thin superconducting film, i.e., the electron pairs are produced between $`(𝐤,n,\sigma )`$ and $`(𝐤,n,\sigma )`$. The direct effective electron-electron interaction due to the exchange of virtual confined slab phonons becomes $$V_{nm}(q)=\underset{l=1}{\overset{l_{max}}{}}M_l^{nm}(q)D_l(q)M_l^{mn},$$ (7) where $`M_l^{nm}`$ is the electron-confined phonon matrix element and $`D_l`$ the phonon propagator , and the $`l`$ sums cover all phonons with energy less than the cutoff energy, $`\omega _D`$. The maximum value of $`l`$ contributing to the sum over the slab phonon mode $`l`$ in Eq. (7), $`l_{max}`$, is given by the condition $`l_{max}=(d/\pi )(\mathrm{}\omega _D/c)`$, where $`c`$ is the velocity of the phonons. (We assume that all slab phonons have the same velocity because information on the slab mode velocity is not experimentally available.) We then get, $$V_{nm}=\frac{J}{Ad}\underset{l=1}{\overset{l_{max}}{}}\left[\beta _{n,n^{}}^{(l)}\right]^2,$$ (8) where the confinement form factor is given by $`\beta _{n,n^{}}^{(l)}`$ $`={\displaystyle \frac{2}{d}}`$ $`{\displaystyle _0^d}𝑑zu_n^{}(z)\mathrm{sin}(l\pi z/d)u_n^{}(z)`$ (9) $`=`$ $`{\displaystyle \frac{2}{\pi }}\left[{\displaystyle \frac{l}{l^2(nn^{})^2}}{\displaystyle \frac{l}{l^2(n+n^{})^2}}\right].`$ (10) Eq. (10) has the following selection rule: for $`|n\pm n^{}|`$ = even (odd) only odd (even) $`l`$’s are allowed in the sum over slab modes $`l`$. The slab phonons with odd $`l`$ are symmetric and the ones with even $`l`$ are antisymmetric with respect to reflection through $`z=d/2`$, i.e., under the transformation $`|zd/2||d/2z|`$. Since the electron wave functions are either symmetric or antisymmetric, the couplings between two subbands of the same symmetry involve symmetric quantized phonons, while couplings between two subbands of different symmetry involve antisymmetric phonons. (Note that this simplicity will be lost if parity is not a good quantum number as it is in our simple infinite square well model, but would not be under an asymmetric confinement.) We see that $`V_{nm}`$ decreases with increasing $`l`$, since the transition of electrons to higher subbands cannot be induced by phonons with small momenta $`q`$. Thus, the components $`V_{nm}`$ of the interaction matrix form a monotonically decreasing sequence with increasing subband index. When all the confined phonons contribute to $`V_{nm}`$ (i.e., $`l_{max}\mathrm{}`$) we recover the bulk phonon mediated results , i.e., $`V_{nm}=(J/Ad)(1+\frac{1}{2}\delta _{nn^{}})`$, as we should. As a result of the confined phonons, a superconducting condensate of Cooper pairs can be produced in a given subband, first because of the attraction due to the electrons in the same subband and second because of transitions from other subbands contributing to the condensate. In the presence of a number of subbands the reduced BCS Hamiltonian of the system is given by $`H`$ $`=`$ $`{\displaystyle \underset{\mathrm{k},n,\sigma }{}}\xi _n(\mathrm{k})c_{\mathrm{k}n\sigma }^{}c_{\mathrm{k}n\sigma }`$ (11) $`+`$ $`{\displaystyle \underset{\mathrm{k},\mathrm{k}^{},\sigma }{}}{\displaystyle \underset{n,m}{}}V_{nm}c_{\mathrm{k}^{}m\sigma }^{}c_{\mathrm{k}^{}m\sigma }^{}c_{\mathrm{k}n\sigma }c_{\mathrm{k}n\sigma },`$ (12) where $`c_{\mathrm{k}n\sigma }^{}`$ is electron creation operator in the n-th subband with spin $`\sigma `$, $`\xi _n(\mathrm{k})=ϵ_n(\mathrm{k})\mu `$ the electron energy in the n-th subband measured from the chemical potential $`\mu `$, and $`V_{nm}`$ the attractive interaction between $`n`$-th subband and $`m`$-th subband mediated by the confined phonons in the film. In the BCS theory (which is what we utilize) the gap function has the same energy cutoff $`\mathrm{}\omega _D`$ as the interaction. In the weak-coupling approximation we have the the superconducting energy gap for $`n`$-th subband given by the gap equation $$\mathrm{\Delta }_n(T)=\underset{m𝐤}{}\frac{V_{nm}}{2V}\frac{\mathrm{\Delta }_m\mathrm{tanh}(E_m/2kT)}{E_m},$$ (13) where $`E_n=(\xi _n^2+\mathrm{\Delta }_n^2)^{1/2}`$. The nondiagonal terms in the sum reflect the possibility of the transition of the electron pair from one subband into another as a result of interaction with confined phonons. Integration over k gives the gap function of the subband $`n`$ at $`T=0`$ K $$\mathrm{\Delta }_n=\frac{Jm}{2\pi }\underset{n^{}}{}\mathrm{sinh}^1\left(\frac{\omega _c}{\mathrm{\Delta }_n^{}}\right)\mathrm{\Delta }_n^{}\alpha _{nn^{}},$$ (14) where $`\alpha _{nn^{}}=V_{nn^{}}/J`$. If all confined phonons contribute equally to the electron-electron interaction, i.e., ($`\alpha _{nm}=1+\delta _{nn^{}}/2`$), then we recover the results of ref. . With the coupled subband interactions Eq. (14) becomes a non-linear coupled subband matrix equation. The critical temperature is given by $$T_c=1.14\omega _D\mathrm{exp}\left(\frac{2\pi d}{Jm_n^{}\alpha _{1n^{}}x_{1n^{}}}\right),$$ (15) where $`x_{1n}=\mathrm{\Delta }_n/\mathrm{\Delta }_1`$, the ratio of the n-th subband energy gap to the lowest subband energy gap. For any given finite width, $`d`$, of the slab, only a finite number of eigenvalues, $`\xi _n`$, contribute; values of $`\xi _n`$ in excess of $`\mu +\mathrm{}\omega _D`$ make vanishing contributions, because then all the $`\epsilon _\mathrm{k}`$’s lie outside the interaction region. Thus, the summation in Eqs. (14) and (15) is only over all the occupied subbands. For a fixed electron density we can find the maximum value of the $`n`$ (or, the highest occupied subband) from the chemical potential. The number density $`n_0=N/V`$ is related to the chemical potential by the relation, $`N=2_{𝐤,n}n_{𝐤,n}`$, where $`n_{𝐤,n}=[\mathrm{exp}(\epsilon _n(𝐤)/k_BT)+1]^1`$ is the Fermi distribution function. At $`T=0K`$ we have $$\mu =(\pi d\mathrm{}^2/\nu M)\left\{n_0+\frac{\pi }{6d^3}\nu (\nu +\frac{1}{2})(\nu +1)\right\},$$ (16) where $`\nu `$ is the maximum value of occupied subband $`n`$ and is given by the integral value of the expression, $`\nu =dk_F/\pi `$. If we write Eq. (15) in the usual BCS form, $`T_c=1.14\omega _D\mathrm{exp}(1/N_{film}J_{eff})`$, the effective interaction parameter can be written as $$J_{eff}=\frac{2J}{2\nu +1}\underset{n=1}{\overset{\nu }{}}\alpha _{1n}x_{1n}.$$ (17) As $`d\mathrm{}`$, $`J_{eff}J`$ because $`\alpha _{1n}(d\mathrm{})=1+\delta _{1n}/2`$ and $`x_{1n}(d\mathrm{})=1`$, and $`N_{film}N_{3D}`$. Thus, we recover the three dimensional results as we expect. In Fig. 1 we show the calculated superconducting energy gaps of each subband and the critical temperature as a function of the film thickness $`d`$ up to the 4-th subband occupation. In this figure we use the following parameters: Debye energy $`\mathrm{}\omega _D=100K`$, electron density $`n_0=2\times 10^{22}cm^3`$, and $`\rho =N_{3D}J=mk_FJ/(2\pi ^2)=0.3`$. In previous calculations the energy gaps of different subbands were find to be the same by virtue of the bulk phonon approximation. The shape resonance feature in the earlier calculations arise only from the effective 2D density of state of the film as the chemical potential passes through different subbands. In our calculation we find that the energy gaps are different for different subbands since the effective coupling strength depends explicitly on the occupied subbands. In addition, the resonance structures in our results (the sharp maxima in Fig. 1) arise from both the thickness dependent density of states $`N_{film}`$ and the effective interaction parameter $`J_{eff}`$. The energy gap is maximum for the lowest subband and decreases in the higher subbands for a fixed film thickness. In Fig. 1(b) we also compare our results with Ref. for equivalent parameter values. Our results exhibit more resonance features and has much lower $`T_c`$ (typically, about half around the maxima) than that of Ref. . Note that any inhomogeneity on the microscopic scale (quite unavoidable in real thin films of 5-15 $`\AA `$ thickness) will considerably suppress the resonance features of Fig. 1, and any enhancement in $`\mathrm{\Delta }_n`$ or $`T_c`$ may remain unobservable unless the films are microscopically of uniform thickness. The reduction of the critical temperature in our calculation can be explained by the enhancement of the effective interaction parameter in the slab phonon model. The present work can easily be extended to include more realistic boundary conditions, but one then needs to resort to numerical work right from the beginning, losing much of the essential qualitative physics of the phenomenon. We believe that the basic physics discussed in this paper and the qualitative features of our results shown in Fig. 1 transcend our specific model, and should be valid in any BCS type superconductivity in thin films. This work is supported by the U.S.-ARO and the U.S.-ONR.
no-problem/9908/hep-th9908061.html
ar5iv
text
# References IMSc/99/08/\** hep-th/9908061 A Note on Holographic Principle in models of Extended Inflation type S. Kalyana Rama Institute of Mathematical Sciences, C. I. T. Campus, Taramani, CHENNAI 600 113, India. email: krama@imsc.ernet.in > ABSTRACT. > > We present a simple derivation of an upper bound on the average size of the true vacuum bubbles at the end of inflation, in models of extended inflation type. The derivation uses the inequality that the total energy inside a given volume must be less than its linear dimensions. The above bound is the same as that obtained earlier, by applying the holographic principle according to Fischler-Susskind prescription. Such a bound leads to a lower bound on the denisty fluctuations. PACS numbers: 98.80.Cq, 98.80.Bp Holographic principle says, among other things, that the entropy in a given volume is bounded by its surface area, measured in planck units. It is based on the inequality $$E<r$$ (1) where $`E`$ is the total energy inside a given volume, and $`r`$ is its linear dimensions. Otherwise the system would lie within the Schwarzschild radius, and would collapse to form a black hole . Recently, Fischler and Susskind (FS) have a given a prescription for applying the holographic principle in cosmology . Applying FS prescription during the inflationary stage, we have obtained in an earlier work a lower bound on the density fluctuations .<sup>1</sup><sup>1</sup>1 FS prescription is said to be invalid during reheating, a non adiabatic process . However, in , the FS prescription is applied only (soon) after the reheating, when the entropy generation is complete. To our knowledge, this is the first instance where a lower bound on density fluctuations has been obtained theoretically. In the explicit cases considered, this bound is remarkably close to the observed value. Later, other prescriptions for applying the holographic principle in cosmology have been suggested . However, none of them seems to lead to a lower bound on the density fluctuations. Hence, the existence of such a theoretical lower bound, important though it is, is in doubt. In this note, we present a simple alternative derivation of this bound, using equation (1). The derivation is applicable to models of extended inflation type, where the inflation is ended by the percolation of true vacuum bubbles (and eventual reheating) . In the following, we use planck units unless mentioned otherwise and, for simplicity, omit numerical factors of order unity. Let $`T_b`$ be the temperature when the inflation begins, and $`\rho _FT_b^4`$ be the false vacuum energy. In extended inflation type models, the true vacuum bubbles nucleate during inflation, expand with the speed of light, and eventually percolate the universe completely, thus ending the inflation. (The bubble walls then collide and reheat the universe to a temperature $`T_RT_b`$.) Let $`d`$ be the average size of the bubbles at the instant of complete percolation and let $`\delta T_b^1`$ be the bubble wall thickness. (Thicker bubble walls will result in stronger bounds.) Then, the energy $`E`$ stored in the bubble is $$E\pi d^2\delta \rho _Fd^2T_b^3.$$ (2) Before percolation is complete, the bubbles are expanding with the speed of light, and are embedded in an inflating universe. It seems unlikely that, under these circumstances, the bubbles can collapse back to form black holes. The inequality (1) may not then be applied. However, upon the completion of percolation, the inflation ends, the bubbles stop expanding, and are now embedded in an expanding universe like ours. Then, regions with sufficient amount of energy inside can collapse to form black holes. The inequality (1) may then be applied. Applying (1) now gives the constraint $$E<ddT_b^3<1.$$ (3) This constraint is same as that in , which was obtained by applying holgraphic principle according to Fischler-Susskind prescription. As explained in with explicit examples, an upper bound on $`d`$ leads to (i) an upper bound on the duration of inflation; (ii) an upper bound on the inflation factor; and (iii) a lower bound on the density fluctuations, $`\frac{\delta \rho }{\rho }`$. Explicitly, in the case of extended inflation with $`T_b10^{14}GeV`$ and $`\omega 10`$, one gets $`\frac{\delta \rho }{\rho }\stackrel{>}{_{}}𝒪(10^7)`$ (see for details). This lower bound, obtained theoretically, is remarkably close to the observed value $`10^6`$ . We conclude with two remarks. First, it is clear from the above derivation that the violation of the bound (3) is likely to result in a copius production of black holes. We are assuming implicitly that such is not the case. On the other hand, it is perhaps of interest to study the consequences, cosmological or otherwise, of such a copious black hole production. However, such a study lies beyond the scope of the present work. Second, it is not clear if, and how, the above derivation can be extended to models of new inflation type, where the inflation is ended by the inflaton ‘rolling down the hill’ .
no-problem/9908/astro-ph9908140.html
ar5iv
text
# References A Cepheid distance to NGC 4258 Eyal Maoz<sup>#,∗</sup>, Jeffrey A. Newman, Laura Ferrarese, Peter B. Stetson, Stephen E. Zepf<sup>§</sup>, Marc Davis, Wendy L. Freedman, and Barry F. Madore<sup>∥,¶</sup> e-mail: maoz@astro.berkeley.edu, jnewman@astro.berkeley.edu, lff@astro.caltech.edu, Peter.Stetson@hia.nrc.ca, zepf@astro.yale.edu, marc@astro.berkeley.edu, wendy@ociw.edu, barry@ipac.caltech.edu <sup>#</sup>NASA Ames Research Center, MS 245-3, Moffett Field, CA 94035-1000 Department of Astronomy, University of California, Berkeley, CA 94720 Hubble Fellow, California Institute of Technology, MS 105-24, Pasadena, CA 91125 Dominion Astrophysical Observatory, 5071 W. Saanich Rd., Victoria, B.C., Canada V8X 4M6 <sup>§</sup>Department of Astronomy, P.O. Box 208101, Yale University, New Haven, CT 06520 Observatories of the Carnegie Institution of Washington, 813 Santa Barbara St., Pasadena, CA 91101 NASA/IPAC Extragalactic Database, Infrared Processing and Analysis Center, Jet Propulsion Laboratory, California Institute of Technology, MS 100-22, Pasadena, CA 91125 Distances measured using Cepheid variable stars have been essential for establishing the cosmological distance scale and the value of the Hubble constant. These stars have remained the primary extragalactic distance indicator since 1929 because of the small observed scatter in the relationship between their pulsation period and luminosity, their large numbers, which allow many independent measures of the distance to a galaxy, and the simplicity of the basic physics underlying their variability. Potential systematic uncertainties in the use of the Cepheid period-luminosity relation to determine distances are estimated to be $`810\%`$$`^{\text{[1]}}`$. Here we describe the results of a search for Cepheids in the nearby galaxy NGC 4258, which has an independently determined geometric distance of 7.2$`\pm `$0.5 Mpc $`^{\text{[2]}}`$. We determine a Cepheid distance of 8.1$`\pm `$0.4 (excluding possible systematic errors affecting Cepheid distances) Mpc; there is a 1.3$`\sigma `$ difference between the measurements. If the maser-based distance is adopted and other Cepheid distances are revised according to our results, the derived value of the Hubble constant would be increased by 12$`\pm `$9%, and the corresponding age of the Universe would decrease by the same factor. Over the last 20 years, substantial improvements have been made in the Cepheid distance scale$`^{\text{[3, 4, 5]}}`$. The Cepheid period-luminosity (P–L) relation derived from observations of stars in the Large Magellanic Cloud (LMC) has formed the basis for the calibration of the HST Key Project on the Extragalactic Distance Scale ($`^{\text{[6]}}`$ henceforward, the Key Project) and of many other recent extragalactic distance measurements$`^{\text{[7]}}`$. Cepheid distances to galaxies as far away as 25 Mpc have been reliably measured using the Hubble Space Telescope (HST) with random errors of a few percent; the results have been used to calibrate a number of secondary distance methods and thereby to provide estimates of the Hubble constant. Today, the largest identified potential sources of systematic error in the Cepheid distance scale are the zero point of the Cepheid period-luminosity relation (or alternatively, the adopted distance to the LMC), difficulties in the calibration of the HST Wide Field and Planetary Camera 2 (WFPC2) camera which has been used for most modern studies, and the possible effects of differences in the chemical composition of stars on Cepheid distance measurements. These have been judged to be $`\pm `$ 6.5%, $`\pm `$ 4.5%, and $`\pm `$ 4% in distance, respectively$`^{\text{[1]}}`$, with substantial non-Gaussianity possible; for instance, recently published measurements of the distance to the LMC range from about 40 to about 55 kpc, with a distribution skewed towards lower values$`^{\text{[8]}}`$. Given the remaining uncertainties affecting the application of the LMC-calibrated Cepheid period-luminosity relation to galaxies observed with HST, further tests of the extragalactic distance scale are clearly important. The spiral galaxy NGC 4258 presents new opportunities to test and potentially to improve the calibration of the Cepheid P–L relation because of the precision with which its distance has been measured in a manner independent of the conventional ladder of astronomical distance scales$`^{\text{[9]}}`$. This distance, $`7.2`$ Mpc, has been determined using its apparently simple, Keplerian circumnuclear disk delineated by line-emitting water masers that orbit a supermassive black hole at its center$`^{\text{[10]},\text{[11]}}`$. The total estimated uncertainty in this distance is $`\pm 0.3`$ Mpc if the disk is presumed to be circular; if nonzero eccentricities are allowed, the uncertainty increases to $`\pm 0.5`$ Mpc (we adopt this value for all further discussion). The direct, geometric methods used are believed to have minimal unknown systematic uncertainties. Combining the observed rotation velocities with the measured centripetal acceleration in the disk$`^{\text{[12]}}`$ or the observed proper motions of the maser sources$`^{\text{[2]}}`$ allows independent measurements of the physical size of the disk; comparing these to its observed angular extent yields the distance to the galaxy center via simple geometry. The two routes to this distance (proper motions and accelerations) yield results in agreement with each other to $`1\%`$. We have therefore observed a portion of NGC 4258 to search for Cepheids using the WFPC2 camera and the Hubble Space Telescope on 11 epochs in 1998 with both the $`F555W`$ and $`F814W`$ filters. Details of the observations and the analysis procedures summarized here will be described elsewhere (Newman et al. 2000, in preparation); full photometric results and other additional figures and tables are also given there. We have performed point-spread-function fitting photometry on these data with two commonly-used software packages – DAOPHOT/ALLFRAME$`^{\text{[13]}}`$ and DoPHOT$`^{\text{[7, 14]}}`$– following Key Project-like procedures throughout$`^{\text{[17]}}`$, and converted results to standard $`V`$ and $`I`$ filter photometric systems. Cepheid searches were performed using two different algorithms with the ALLFRAME dataset$`^{\text{[15, 16]}}`$, and using a third for DoPHOT data$`^{\text{[17]}}`$. We have identified and determined light curves, periods, mean magnitudes, and colours for 15 Cepheids. All of these stars fulfill four criteria: they are identified as variable by all three search techniques, fit a Cepheid template light curve with reasonable $`\chi ^2`$, visibly vary in blink comparisons in both $`F555W`$ and $`F814W`$ images, and have negligible statistical probability of being misidentified nonvariables. Light curves illustrating the variability of several of the Cepheids found are plotted in Figure 1. The DoPHOT NGC 4258 P–L relations are plotted in Figure 2. We have adopted the DoPHOT photometry for all major conclusions reported here, as ALLFRAME photometry yielded internally discrepant distances from Cepheids on the two WFPC2 chips used (by 0.24 mag, a 1.8$`\sigma `$ difference possibly due to an error in aperture corrections), whereas the corresponding DoPHOT results agreed to within 0.07 mag (0.5$`\sigma `$); however, as an additional check we also provide ALLFRAME values in much of what follows. In Key Project procedures, the differences between the observed and intrinsic colours of Cepheids are then used to correct for the effects of extinction by dust assuming a standard Galactic reddening law$`^{\text{[18]}}`$. Applying this correction star-by-star and then robustly averaging yields results for small datasets that best match those that would be obtained using Key Project methods with more Cepheids (Newman et al. 2000, in preparation), so we adopt this as our primary method of determining a distance modulus for NGC 4258. This technique yields a distance with a statistical uncertainty (derived from the P–L relation fit) of $`\pm 0.07`$ mag. Based upon past experience with Key Project galaxies, we would expect a possible 1$`\sigma `$ systematic error of $`\pm 0.04`$ mag from the uncertainties in photometry due to the difficulty of obtaining accurate aperture corrections in our fields, which contained relatively few bright, isolated stars. This error should be highly correlated between $`V`$ and $`I`$ measurements and thus not propagated through the reddening correction. However, to be conservative we instead adopt the entire ALLFRAME-DoPHOT distance modulus difference (0.10 mag) as an estimate of possible errors affecting only photometry for NGC 4258. Adding this in quadrature to the random error, the total uncertainty unique to our determination of a Cepheid distance to NGC 4258 is $`\pm 0.12`$ mag. This measurement is subject to a number of potential sources of systematic error that also affect Key Project distance determinations, as described in Table 1; their possible contributions have been estimated to total $`\pm 0.18`$ mag$`^{\text{[19]}}`$. Further description of these sources of error may be found in Key Project papers$`^{\text{[30]}}`$. Correcting for differences in heavy element content between the field we have studied and the LMC$`^{\text{[20]}}`$ using previously published studies of NGC 4258$`^{\text{[21]}}`$ would lead to an increase of $`0.08\pm 0.06`$ mag in the distance modulus we have derived. However, due to the lack of agreement on the magnitude and sign of this effect in recent studies, the Key Project has refrained from applying such corrections but considered the resulting uncertainty to be a possible systematic error. We thus do likewise, and obtain a Cepheid distance modulus to NGC 4258 of $`29.54\pm 0.12`$ mag (unique to this determination) $`\pm 0.18`$ mag (systematic uncertainties in Key Project distances), corresponding to a metric distance of 8.1 $`\pm 0.4`$ Mpc $`\pm 0.7`$ Mpc. When treated in the same way, the ALLFRAME results yield a distance modulus of $`29.64\pm 0.09`$ mag, corresponding to a metric distance of $`8.5\pm 0.5`$ Mpc. The distance to NGC 4258 derived from observations of Cepheids is thus appreciably greater than the maser distance of 7.2 $`\pm 0.5`$ Mpc$`^{\text{[2]}}`$. The Cepheid and maser distances differ by 1.0$`\sigma `$ if we add our measurement uncertainty of 0.4 Mpc, the Key Project systematic error estimate of 0.7 Mpc, and the maser distance error estimate of 0.5 Mpc in quadrature; potential systematic errors in either technique do not seem to have been grossly underestimated. If we assume that the maser distance is correct (up to its stated uncertainty) and wish to determine whether a revision of the Cepheid distance scale may be desirable, however, prior estimates of systematic errors in the Cepheid distance scale are irrelevant to calculations of significance. In that case, there is a 1.3$`\sigma `$ (80% significance) discrepancy (1.6$`\sigma `$ if the masing disk is presumed to be circular). The largest potential source of such a difference in either error budget is the uncertainty in the distance modulus to the LMC; however, other identified potential sources of systematic error are great enough that a revision of the LMC distance is not required by our results. A number of other tests of the extragalactic distance scale have been attempted in recent years using independent, physical or geometric methods; however, none of them currently have the same precision and direct applicability to recent Cepheid measurements that the maser distance to NGC 4258 may provide in combination with our work. Observations of a “light echo” from Supernova 1987A have been used to place upper limits on the distance to the LMC; however, results differing by 10% in distance from each other have been obtained with this method$`^{\text{[22, 23]}}`$, and it cannot place constraints on errors related to WFPC2 calibration or the effects of differences in chemical composition from the LMC. Interferometric radio observations of the expansion of SN1993J in M81 have been used to measure its distance with $`15\%`$ uncertainty under modest assumptions $`^{\text{[24]}}`$; the agreement with the HST Cepheid distance to its host galaxy is excellent$`^{\text{[25]}}`$. However, the uncertainty in the distance to this supernova exceeds the estimated systematic uncertainties in the Cepheid distance scale, and the Cepheid distance to M81 was obtained with a different instrument (WF/PC) and photometric calibration techniques than were used for the majority of Key Project galaxies. Finally, distances to several Key Project galaxies have been obtained using the Expanding Photospheres Method on supernovae they hosted, again with good agreement. However, all of those supernovae were abnormal and considered a priori to be poor candidates for the method used; therefore, there may be substantial systematic errors in those EPM measurements, and drawing strong conclusions from this agreement may be questionable$`^{\text{[26]}}`$. Under the assumption that the maser distance and the estimates of its uncertainty are correct, we can derive the resulting absolute magnitudes of the Cepheids observed and obtain a potential recalibration of the Cepheid P–L relation; the details of such an analysis will be described in another paper (Newman et al. 2000, in preparation). Such a calibration could be significantly less subject to systematic effects than one based on ground-based observations of the LMC, as Cepheids in NGC 4258 should have a chemical composition similar to those in galaxies used by the Key Project to calibrate other distance indicators and have been observed with the same instruments, filters, and parameter measurement techniques used in other HST studies. To lowest order, extinction-corrected Key Project-like HST WFPC2 Cepheid results would be revised using the NGC 4258 maser distance simply by subtracting $`0.25\pm 0.19`$ mag from their distance moduli ($`0.33\pm 0.20`$ mag if a correction for differences in heavy element content from the LMC were made to the NGC 4258 Cepheid distance). Because the chemical composition of the field studied in NGC 4258 matches the average for the Key Project galaxies used to calibrate most secondary indicators, if distances to both are corrected for differences from the LMC and then revised so that the Cepheid distance to NGC 4258 would match the maser distance, the net correction would remain 0.25 mag. The results for NGC 4258 thus would imply that typical HST WFPC2 Cepheid distances may be too high by $`12\%\pm 9\%`$ ($`18\%\pm 9\%`$ if the ALLFRAME results were adopted). It may be worthwhile to note that if applied to all WFPC2 Cepheid distances, such a correction would eliminate the discrepancies between the Tully-Fisher relation calibrated using ground-based distances and that obtained using Key Project results$`^{\text{[27]}}`$, reducing its overall scatter accordingly. If such a comparison is fair, this suggests that the difference in the maser and Cepheid distances may be more likely to be due to the difficulty of WFPC2 calibration than to an error in the assumed LMC distance. However, selection effects or systematic errors in measuring the total galaxy magnitudes of the ground-based calibrators, which are relatively large in angular extent, could also explain the Tully-Fisher discrepancy, as those magnitudes are required for determining such distances. Revising Cepheid-calibrated distance indicators according to the maser distance would increase the measured Hubble constant by $`12\%\pm 9\%`$, and decrease the corresponding age of the Universe similarly. The resulting values would be increasingly difficult to reconcile with globular cluster ages unless the Universe has very low density or, particularly, if it has a nonnegligible cosmological constant or similar negative-pressure component$`^{\text{[28]}}`$. However, the statistical significance of such a revision remains limited. A more compelling test of the HST Cepheid distance scale based on the maser distance to NGC 4258 would require a substantially larger sample of Cepheids (reducing the uncertainties in determining the $`VI`$ P–L relations and the reddening) and better determination of aperture corrections; these issues can be addressed simultaneously by searching for Cepheids with HST in a field that contains more stars and has undergone more recent star formation. It would be reasonable to expect that observations of a region richer in Cepheids might yield 3 times as many Cepheids (giving a distance modulus uncertainty of 0.04 magnitude) and aperture corrections accurate to $`\pm 0.04`$ magnitude; better agreement between ALLFRAME and DoPHOT analyses might also occur with improved data. Reductions of uncertainties in the maser distance, e.g. via improved constraints on the eccentricity of the circumnuclear disk, would also be greatly beneficial for its use to calibrate the extragalactic distance scale. Successful maser distances to other galaxies, establishing a Hubble relation, would more firmly establish this novel technique. With improvements in both Cepheid and maser analyses, NGC 4258 has great potential for establishing a new primary step in the distance ladder, reducing the potential systematic errors in measurements of the Hubble constant to perhaps as little as 5%. Figure and Table Captions Figure 1 – Light curves for three representative examples of the Cepheids we have discovered in NGC 4258. Open symbols depict the measured magnitudes of a given star, and the error bars the uncertainty therein; the solid lines show template Cepheid light curves derived from observations of stars in the LMC$`^{\text{[29]}}`$ that have been used to fit the data. The left panel for each star shows the variation of the $`V`$ magnitude of each star versus the the phase of the variation, which makes the Cepheid pulsation pattern clear. The right panel plots the $`I`$ variation of the Cepheids in similar fashion. Figure 2 – NGC 4258 Period–Luminosity relations. a, The $`V`$ P–L relation for the Cepheids we have found in NGC 4258, based upon DoPHOT photometry. The Cepheid P–L relation has previously been found to be well fit by a linear relationship between mean magnitude and the logarithm of the period in the range 10–60 days. To minimize the effects of incompleteness and to optimize results given the limited number of Cepheids observed, we adopt LMC P–L relation slopes$`^{\text{[3]}}`$ and only fit for differences in the zero point. Open symbols depict the measured parameters of NGC 4258 Cepheids; the solid line indicates the fit P–L relation, and the dashed lines the 2$`\sigma `$ observed scatter of fiducial LMC Cepheids about its P–L relation. From the difference between the absolute magnitudes of LMC Cepheids (for an assumed LMC distance of 50 kpc) and the observed magnitudes of NGC 4258 Cepheids we derive a $`V`$ distance modulus, defined by $`\mu =5\mathrm{l}\mathrm{o}\mathrm{g}_{10}d/10\mathrm{p}\mathrm{c}`$, of $`\mu _V`$=$`29.82\pm 0.07`$ (ALLFRAME results yield $`29.88\pm 0.08`$). This value must be corrected for the effects of dust extinction to yield a reliable measurement of the distance of NGC 4258. b, The $`I`$ P–L relation for NGC 4258 Cepheids. Applying procedures similar to those for $`V`$, we derive an $`I`$ distance modulus $`\mu _I`$=$`29.71\pm 0.05`$ (ALLFRAME photometry yields $`29.78\pm 0.05`$). Correcting for the effects of dust on the apparent magnitude of each star using its observed colour, we obtain an extinction-corrected distance modulus of $`29.54\pm 0.07`$ mag from DoPHOT photometry for all Cepheids; those on WFPC2 chip 2 alone yield $`29.58\pm 0.12`$, while those on chip 3 yield $`29.51\pm 0.09`$ (for ALLFRAME photometry, the corresponding numbers are $`29.64\pm 0.08`$, $`29.50\pm 0.12`$, and $`29.74\pm 0.10`$ mag). Table 1 – A summary of the major identified errors which may affect our extinction-corrected distance modulus measurement. All numbers listed are in magnitudes; for small differences, the corresponding fractional error in distance may be obtained by multiplying by 0.46. For those potential systematic errors which affect all Cepheid distances obtained in the same manner as ours uniformly, we have adopted the uncertainty estimates of the Key Project$`^{\text{[30]}}`$; more detailed descriptions may be found therein. The estimated systematic errors in WFPC2 photometry (S<sub>2</sub>) include uncertainties in zero points and the ‘long versus short’ uncertainty, added in quadrature. Because the metallicity of the field we have studied in NGC 4258 matches the average metallicity of Key Project galaxies used to calibrate the Tully-Fisher relation, surface brightness fluctuation distances, and the peak luminosities of Type Ia supernovae, we treat the uncertainties in metallicity correction as a systematic error common to this distance measurement and those of the Key Project (S<sub>3</sub>). The random uncertainty in the Cepheid distance modulus (R<sub>1</sub>) has been estimated from the standard error of the extinction-corrected moduli for individual stars. We have adopted a conservative estimate of the total photometric uncertainties unique to NGC 4258 (S<sub>4</sub>) of 0.10 mag based on the difference between the overall distance moduli they yield. For chip 2, the mean difference between ALLFRAME and DoPHOT magnitudes for 24 bright, isolated stars was $`0.026\pm 0.049`$ (standard deviation) mag for $`V`$, and $`0.015\pm 0.049`$ mag for $`I`$. For chip 3, the mean difference for 30 stars was $`0.025\pm 0.027`$ mag for $`V`$, and $`0.088\pm 0.046`$ mag for $`I`$. Mean magnitudes for Cepheids yielded results consistent with these to within 1$`\sigma `$, albeit with much larger standard errors. We consider the chip 3/$`I`$ results to be an aberration closely related to the discrepant distance moduli obtained with ALLFRAME on the two chips; even including it, the mean ALLFRAME-DoPHOT magnitude difference among the 4 cases would be 0.04 mag.
no-problem/9908/quant-ph9908077.html
ar5iv
text
# Quantum Consciousness ## 1 Introduction For conscious states and brain states to mirror one another in any species, thereby establishing what von Neumann calls a psycho-physical parallelism, these intrinsically different states must evolve together and interact with one other during their time of evolution. Standard physics makes no provision for an interaction of this kind, but a quantum mechanical opening for an objective/subjective interaction is shown to exist, and is described in previous papers. Our theory of subjective evolution calls for the existence of a Central Mechanism ($`CM`$) within an evolving organism, which contains presently unknown components of the nervous system. The function of a $`CM`$ is to reduce quantum mechanical superpositions within the nervous system, and to simultaneously give rise to a conscious experience of the eigenvalues of the reduction. This accords with von Neumann s requirement that a quantum mechanical state reduction is accompanied by an observer’s conscious experience of the measured variables. At the present time, no one knows what there is about a conscious organism that gives rise to either consciousness or state reduction. We simply combined these two mysteries inside the $`CM`$, thereby placing our ignorance in a black-box so we can ask another question, namely: how do physical and mental states evolve interactively to insure the psycho-physical parallelism? The model in references 2 and 3 requires that a conscious organism spontaneously creates a profusion of macroscopic quantum mechanical superpositions consisting of different neurological configurations. A mechanism for this generation is proposed by H. Stapp. The result is a superposition of different neurological states, each of which may be accompanied by a different subjective experience. A reduction to a single eigenstate is not assumed to be triggered microscopically along the lines of Ghirardi-Rimini-Weber; but rather, it is assumed to occur in response to a macroscopic event. It occurs the moment an emerging subjective state becomes actively conscious in one of the macroscopic neurological components of a Stapp superposition. The consciousness that is associated with such a reduction is assumed to fade the moment reduction is complete, and the resulting subjective *pulse* is supposedly followed by similar pulses in rapid succession. This can make the subject aware of an apparent continuum of consciousness. Presumably, any reduction of this kind is accompanied by a reduction of all other parts of the organism as well as all those parts of the external world that are correlated with it. This means that a second observer, coming on the heels of the first, will make an observation in agreement with the first. More formally, a measurement interaction establishes correlations between the eigenstates $`|a_i`$ of some apparatus (with discrete variables $`a_i`$), eigenstates of a first observer$`|\mathrm{\Phi }_i`$, and eigenstates of a second observer $`|\mathrm{\Theta }_i`$, such that the total state prior to reduction is given by $`|\mathrm{\Psi }=\mathrm{\Sigma }_iC_i|a_i|\mathrm{\Phi }_i|\mathrm{\Theta }_i.`$ The coefficient $`C_i`$ is the probability amplitude that the apparatus is in state $`|a_i`$. Let the first observer become consciously aware of the apparatus variable $`a_k`$. The resulting reduction is a projection in Hilbert space that is found by applying the projection operator of that observer $`|\mathrm{\Phi }_k`$$`\mathrm{\Phi }_k|`$ to the total state. | $`|\mathrm{\Phi }_k\mathrm{\Phi }_k||\mathrm{\Psi }=C_k|a_k|\mathrm{\Phi }_k|\mathrm{\Theta }_k`$ | (1st reduction) | | --- | --- | Let the second observer then become consciously aware of the apparatus variable $`a_m`$. The subsequent reduction is found by applying the projection operator of that observer $`|\mathrm{\Phi }_m`$$`\mathrm{\Phi }_m|`$ to the first reduction. | $`|\mathrm{\Theta }_m`$$`\mathrm{\Theta }_m|`$$`C_k`$$`|a_k`$$`|\mathrm{\Phi }_k`$$`|\mathrm{\Theta }_k`$=$`\delta _{km}C_k`$$`|a_k`$$`|\mathrm{\Phi }_k`$$`|\mathrm{\Theta }_m`$ | (2nd reduction) | | --- | --- | Only if $`m=k`$ is the probability non-zero that the second observer will make a measurement. The second observer therefore confirms the results of the first observer that the apparatus has been left in the eigenstate $`|a_k`$. Again, many of the particulars of a reduction (such as its nonlinearly) are ignored in this paper so we can concentrate on the influence of subjective states on physiological states. To this end we require that *when the emerging subjective states of a neurological superposition are different from one another, they will generally exert an influence on their relative probability amplitudes that is a function of that difference.* In particular, we imagine that when a painful subjective state emerges in superposition with a pleasurable subjective state, the probability amplitude of the painful state will be decreased relative to the probability amplitude of the pleasurable state. No currently known observation contradicts this conjecture, for no previously reported experiment deals specifically with the creation of different observers experiencing different degrees of pain, arising on different components of a quantum mechanical superposition. Let $`N`$ in fig. 1 represent the nervous system of the first primitive organism that makes a successful use of the subjective experience of pain . In a previous paper we imagine this creature to be a fish. It is supposed that the fish makes contact with an electric probe, at which time its nervous system splits into a superposition ($`via`$ the Stapp mechanism) consisting of a withdrawal behavior $`W`$ that is accompanied by \[no pain\], and a continued contact behavior $`C`$ that is accompanied by \[pain\]. The probability of survival of each component in this highly artificial model is initially assumed to be 0.5. However, because of the hypothetical influence of subjective pain on probability amplitudes, only the withdrawal state is assumed to survive the reduction in this idealized example. State reduction in fig. 1 is represented by the horizontal arrow. If $`W`$ is further-more a good survival strategy from the point of view of evolution, then the association $`W`$\[no pain\] and $`C`$\[pain\] will serve the species well, whereas a wrong association $`W`$\[pain\] and $`C`$\[no pain\] will lead to its demise. It does not matter to the above argument if the variables are pleasure/pain or some other range of subjective experiences. If a subjective experience like ‘A’ increases the probability amplitude of an escape behavior, and if a subjective experience like ‘B’ diminishes the probability amplitude of that behavior, and if the escape is one that moves the creature away from something that is dangerous to its health, then a distant descendent will experience ‘A’ associated with life supporting escapes, and ‘B’ associated with life threatening failures-to-escape. It is apparent that the quality of the experience does not matter. We require only that the subjective experience in question has a predictable plus or minus effect on the probability amplitudes within a superposition, and the survival mechanisms of evolution will do the rest. They will insure that the eventual subjective life of a surviving species mirrors its experiences in a definite and predictable way thereby establishing a reliable psycho-physical parallelism. We assume that ordinary $`perception`$ do not have this effect. They do not give rise to the hypothetical feedback. In fig. 2 we imagine the existence of an externally imposed two component superposition consisting of environments $`e_1`$ and $`e_2`$, which is produced by using, say, a $`\beta `$ source. The two environments are assumed to have equal probability, and are allowed to interact with the subject s nervous system given by $`N_0`$. Before a reduction can occur, two conscious states emerge from the interaction represented by the superposition of $`(eN)_1[x_1]`$ and $`(eN)_2[x_2]`$, where the conscious part shown in brackets is the observed eigenvalue $`x`$ associated with components 1 and 2. Since we require that an observer of the perceived variable $`x`$ cannot affect the probability of $`x`$, the pure state reduces to a mixture having the same probability as the initial superposition (horizontal arrows in fig. 2). State $`e_i`$ represents the relevant laboratory apparatus together with the wider environment with which it is entangled. The phase angles $`\varphi `$ and $`\varphi ^{}`$ are definite, but they are not localized to manageable parts of the apparatus.. We call them ”arbitrary” in this paper to indicate that their values are not practically calculable, and to emphasize the lack of coherence between these ”macroscopic” components. On the other hand, if pain were the variable in fig. 2 rather than the externally perceived variable $`x`$, it is suggested by our hypothesis that the resulting mixture might no longer be a 50 - 50 split. This possibility is represented in fig. 3, where the final mixture probabilities are left unspecified because they must be discovered by observation. The author has now performed an experiment of this kind with the result reported below. ## 2 The Experiment Two scalers L and R recording local background radiation are placed side-by-side in fig. 4. Their outputs are fed to a selector box that chooses channels L or R, depending on which is the first to record a single count after the selector has been turned on. A 20 V signal is then emitted from the output of the chosen channel. The output on the R-channel is unused, but the L-output closes a relay that puts 80 volts across two metal bars. Two seconds after the selection, an L or R-light goes on indicating which channel was selected. A finger placed across the metal bars will receive a painful 80V shock when the L-channel is selected. This apparatus allows us to carry out the experiments diagramed in figs. 2 and 3. If the selector is initiated in the absence of an observer, we say that the system will become a macroscopic superposition given by $`(e^{i\varphi }e_1+e_2)`$, where $`e_1`$ is the entire apparatus following an L-channel activation, and $`e_2`$ is the entire apparatus following an R-channel activation. The incoherence of the two components (represented by the arbitrary angle $`\varphi `$) is generally understood to mean that the system is indistinguishable from a classical mixture, since interference between these macroscopic components is not possible. However, for reasons given in previous papers, we claim that the final state is really an incoherent quantum mechanical superposition rather than a classical mixture.<sup>2</sup><sup>2</sup>2The uncertainty associated with a classical mixture state represents an outsider’s ignorance, whereas a pure quantum mechanical state superposition represents an uncertainty that is intrinsic to the system (see ref. 1, pp. 1622, 1624; and ref. 2, bottom of p. 1703). Following von Neumann, we assume that the initial intrinsic uncertainty (concerning which of the scalers fires first) will remain an intrinsic uncertainty until it is reduced by “observation”. Hence, the apparatus will remain a macroscopic pure state quantum mechanical superposition until an observation occurs. The lack of interference between the components has no bearing on our result because the hypothetical effect described in this paper relates to, and directly affects, probability amplitudes only. The effect we are looking for should be observable with or without coherence between L and R. If an observer is present and exposed only to the L-light or the R-light, then a reduction will occur like the one in fig. 2, where eigenvalues $`x_1`$ and $`x_2`$ represent a conscious experience of one or the other of those lights. If the observer is exposed only to a conscious experience of “pain or no pain” through his finger across the metal bars, then a reduction like the one in fig. 3 will occur. This experiment may not appear to be quantum mechanical, but it is quantum mechanical by virtue of the particular hypothesis that is being tested in fig. 3. The equipment in fig. 4 was used for a total of 2500 trials, each consisting of two parts. The author s finger was first placed across the metal bars, the selector was turned on, and a shock or no shock was recorded before the lights were observed. In the second part of each trial the finger was replaced by an equivalent resistance, the selector was again initiated, and the appearance of the L or R channel light was recorded. Total number of trials . . . . . . . . . . . . . . . . . . . . . . . . . . $`N`$ = 2500 Number of shocks received in the first part . . . . . . . . . . . . . $`N_S`$ = 1244 Number of times the L-light went on in the second part . . . . . $`N_L`$ = 1261 There are three possible outcomes of a single trial. Either the difference $`N_L`$ \- $`N_S`$ increases, or it decreases, or it remains the same. The three possibilities are represented by the variables $`u`$ (increase) occurring with a probability $`p`$, and $`d`$(decrease) with a probability $`q`$, and $`e`$ (remain the same) with a probability $`r`$. It was found in the experiment that $`u`$ = 632 and $`d`$ = 615 after 2500 trials. If we approximate $`p_0=N_L/N`$ to be the probability that the left channel fires in the second part of each trial (absent the finger), and $`q_0=1p_0`$ to be the probability that the right channel fires in the second part of each trial, then | $`p_0=1261/2500=0.5044`$ | $`q_0=0.4956`$ | | --- | --- | Assuming as a null hypothesis that there is no statistical difference between the displacement of a finger across the metal bars and an equivalent resistor, we have $`p=p_0q_0`$, $`q=q_0p_0`$, and $`r=p_0^2+q_0^2`$, giving | $`p`$ = 0.2500 | $`q`$ = 0.2500 | $`r`$ = 0.5000 | | --- | --- | --- | The variances of $`(u+d)`$ and $`(ud)`$ are $`\sigma ^2(u+d)=<(u+d)^2><u+d>^2=\sigma ^2(u)+\sigma ^2(d)+X`$ $`\sigma ^2(ud)=<(ud)^2><ud>^2=\sigma ^2(u)+\sigma ^2(d)X`$ therefore $`\sigma ^2(ud)=2\sigma ^2(u)+2\sigma ^2(d)\sigma ^2(u+d)`$ $`=2p(q+r)N+2q(p+r)Nr(p+q)N`$ or $`\sigma (ud)=[[4pq+r(p+q)]N]^{1/2}=[N/2]^{1/2}=35.4`$ Our alternative hypothesis is that $`ud`$ is significantly different from 0. But from the data, $`ud=N_LN_S=17`$ after 2500 trials, and this is well within the above the standard deviation around 0. The separate variables $`u`$ and $`d`$ are also within the standard deviation $`\sigma (u)=\sigma (d)=[p(q+r)N]^{1/2}=21.7`$ of their expected value of 625. One can always argue that the statistics are inadequate to reveal a significant difference between $`u`$ and $`d`$. However, they are sufficient to convince the author that the presence of pain on one component of this externally imposed superposition has no significant effect on the outcome. We therefore conclude that the reduction in fig. 3 is not affected by the subjective content of the square brackets in that figure. Further details about this experiment can be found in the document ”QC Experiment” on the author s home page. ## 3 Model Modification This result forces us to make a distinction between externally imposed superpositions and the superpositions created internally by a $`CM`$. In fig. 5 we let a $`CM`$ interact with an environmental superposition, where the latter includes all that is not contained in the $`CM`$ (including other possible $`CM`$s within the organism). The first pair of diverging arrows in that figure carries the initial product into components $`(eCM)_1`$ and $`(eCM)_2`$. This reaction goes according to Schrö dinger. The subsequent pairs of diverging arrows shown in fig. 5 denote the appearance of new superpositions that are produced by the (Schrö dinger) process proposed by Stapp. It is at this point that we engage the (non-Schrö dinger) hypothesis of sect. 1 that allows the emerging ‘subjective’ states on different components of these $`CM`$-superpositions to vary their own probability amplitudes. However, we add the further stipulation that *any such variation can only occur relative to the other components of the originating $`CM`$-superposition.* This restricts the range of states over which the relative amplitude variation can take place. It is thereby required that any variation (due to the hypothetical subjective influence) that takes place within $`(eCM)_1`$ in fig 5 has no effect on a variation within $`(eCM)_2`$, and vice versa. More generally, the probability amplitude of any $`CM`$ as a whole is not affected by a variation that takes place outside of itself, so the normalization of each component $`(eCM)_i`$ of the superposition is preserved. This modification is consistent with the results of the experiment in sect. 2. The superposition in our experiment is represented by the first pair of diverging arrows. Since it is external and not created by the *CM*, it cannot, according to the above stipulation, be affected by the subjective content of components 1 and 2. On the other hand, the second pairs of diverging arrows represent superpositions that *are* created by the *CM*s 1 and 2, so they are subject to our hypothesis of sect. 1. On this modified model, subjective evolution does not rely in any way on externally created superpositions, thereby insuring that the observer of (external) quantum mechanical systems will always record eigenvalues with the probability predicted by standard theory. At the same time, our hypothetical subjective influence can still be realized within either one of the $`CM`$-superpositions in the figure. With this modification, the hypothesis becomes much more difficult to demonstrate experimentally. Although verification remains possible in principle, it will require a more detailed understanding of the workings of the nervous system and/or other parts of the body. ## 4 Bio-Active Peptides Neurological communication depends on the diffusion of chemical neurotransmitters across the synaptic junction between neurons. There is another communication system within the body that makes use of chemicals that are produced at one site and received at another; but in this case, the distances between a production and receiver sites are macroscopic. About 95% of these chemical communicators are peptides, which are mini-proteins consisting of up to 100 amino acids having a maximum atomic mass of 10,000 u. Their classical dimensions are $`\mathrm{\Delta }`$x =10 nm at most, which we assume approximates their size close to the production site. Therefore, Heisenberg tells us that the minimum quantum mechanical uncertainty in the velocity of one of these free peptides is $`\mathrm{\Delta }`$v =0.63 mm/s. Peptides are carried through intercellular space by blood and cerebrospinal fluid. They do not move very far in a tenth of a second, but in that time the Heisenberg uncertainty in position of a peptide will be at least $`\mathrm{\Delta }`$s = $`\mathrm{\Delta }`$v$`\mathrm{\Delta }`$t = 63 mm. This is an enormous uncertainty of position relative to one of the peptide receptor sites which has a size similar to that of the peptide, and which is often separated from its neighbors by comparable distances. Therefore, quantum mechanical uncertainty is an important factor in determining the probability that a given peptide is captured by a given receptor. Stapp s mechanism for introducing quantum mechanical superpositions into the brain relies on the uncertainty in the position of calcium ions in neuron synapses. We suggest that peptides represent another possible source of super-positions that may be just as widespread. And because peptides play an important role in the chemistry of the body, they too may have a significant quantum mechanical influence on behavior. As with the Stapp mechanism, one might object that the uncertainty associated with the peptide s classical diffusion during its migration will overwhelm the quantum mechanical uncertainty, or that a large number of migrating molecules will obscure all quantum mechanical effects. However, the classical uncertainty associated with many-particle ensembles has only to do with our ignorance of initial conditions. In reality, the only uncertainties a receptor will see are those associated with an incoherent quantum mechanical superposition of pure peptide states. This superposition will have as many components as there are peptide molecules involved. And since our hypothetical influence acts through the amplitude of these components, the presence of a large number of independent particles will only increase the hypothetical influence. ## 5 Drugs There are many drugs that can be introduced into the body that will compete with endogenous peptides to occupy the body s receptor sites, and some of these drug molecules are small enough to have a very large quantum mechanical uncertainty of position. For this reason, peptide/drug superpositions are more promising for the purpose of experimental manipulation than calcium ion super-positions. For example, $`endorphins`$ are peptides that unite with special receptors to eliminate pain and/or produce euphoria. They and their receptors can be found everywhere in the body, but they are most intensely located in the limbic system of the brain. There is a drug called $`naloxone`$ that is a strong competitor with the endorphins to occupy the same receptors, and it has the property that it reverses the analgesic/pleasurable effects of the endorphins. If endorphin molecules and externally administered naloxone molecules are in quantum mechanical superposition with one another as their sizes and likely time together suggests, and if they both compete with one another for successful attachment to the same receptor site, then the ratio of endorphin attachments to naloxone attachments would (according to our hypothesis) be a function of the competing subjective states. Since the difference in subjective effects between these two molecules is considerable along the pleasure/pain spectrum, an experimental design involving endorphin/naloxone superpositions appears to offer an opportunity to test the modified model proposed in sect. 3. The author is not able to propose a specific experiment at this time, but an approach along these lines seems promising. ## 6 Evolutionary Advantage It was pointed out in a previous paper that our evolutionary mechanism of objective-subjective interaction (represented by fig. 1) does not insure that a creature evolving under its influence will evolve more quickly or be more successful than a creature evolving strictly as an automaton. That will be true as well of the modified model in sects. 3-5. However, it is not unreasonable to suppose that both conscious evolution and autonomic evolution might work separately and in tandem with one another. The kinds of neurological changes that are necessary for autonomic evolution might very well be independent of the kinds of neurological changes that are necessary for quantum/consciousness evolution. If that is so, and if these two processes work in tandem, then the evolution of the organism will be faster than either the autonomic route by itself, or the conscious route by itself. One would then be able to say that the introduction of consciousness as proposed in this paper will always work to the advantage of the organism. ## Acknowledgement The author would like to thank Erlend Graf for his many valuable comments and suggestions.
no-problem/9908/nucl-th9908026.html
ar5iv
text
# Secondary Charmonium Production in Heavy Ion Collisions at LHC Energy ## 1 Secondary charmonium production The initial energy density in ultrarelativistic heavy ion collisions at LHC energy exceeds by a few order of magnitudes the critical value required for quark-gluon plasma formation. Thus, according to Matsui and Satz , one expects the formation of charmonium bound states to be severely supressed due to Debye screening. The initially produced $`c\overline{c}`$ pairs in hard parton scattering, however, due to charm conservation, will survive in the deconfined medium until the system reaches the critical temperature where the charm quarks hadronize forming predominatly $`D`$ and $`\overline{D}`$ mesons. The appreciable number of $`c\overline{c}`$ pairs and consequently $`D`$,$`\overline{D}`$ mesons expected in Pb-Pb collisions at LHC energy can lead to additional production of charmonium bound states due to the reactions such as, $`D\overline{D^{}}+D^{}\overline{D}+D^{}\overline{D^{}}\psi +\pi `$ and $`D^{}\overline{D^{}}+D\overline{D}\psi +\rho `$ as first indicated in . In this work we present a quantitative description of secondary $`J/\psi `$ and $`\psi ^,`$ production due to the above processes from the thermal hadronic medium created in Pb-Pb collisions at LHC energy. ## 2 Cross section and production rate The charmonium production cross section $`\sigma _{D\overline{D}\psi h}`$ can be related to the hadronic absorption of charmonium $`\sigma _{\psi hD\overline{D}}`$, through detailed balance. The magnitude of charmonium absorption cross section on hadrons is still, however, theoretically not well under control. In fig.1 we show the predictions for the dissociation cross section of $`J/\psi `$ on pions of: (1) the quark exchange model , (2) the comover model with $`\sigma _{\psi \pi }`$3mb , (3) calculations using an effective hadronic Lagrangian and (4) a short distance QCD approach . The large theoretical uncertainties of the cross section seen in fig.1 will naturally influence the yield of secondary charmonium bound states. In the thermal hadronic medium the rate of charmonium production from $`D\overline{D}`$ annihilation is determined by the thermal average of the cross section and the densities of incoming and outgoing particles . The solution of the rate equation requires additional assumptions on the space-time evolution of the hadronic medium and the initial number of $`D`$ and $`\overline{D}`$ mesons at the beginning of the mixed phase. We have adopted a hydrodynamical model for expansion dynamics assuming that at initial time $`\tau _00.1`$fm the system is created as an equilibrium quark-gluon plasma of temperature $`T_01`$GeV and is then undergoing isentropic longitudinal expansion with a superimposed transverse flow. In fig.2 we show the time evolution of the temperature in our model for two values of $`T_c`$ with 12000 pions (charged and neutral)per unit of rapidity in the final state. In the actual numerical calculation we take $`T_c0.17`$GeV as value for the critical temperature, following recent calculations within the lattice gauge theory approach . The analysis of presently available data for the yield of hadrons produced in heavy ion collisions suggests that the chemical freezeout temperature at LHC should be in the range $`0.16<T_f<0.17`$ , i.e very close to $`T_c`$. To get the initial number of $`c\overline{c}`$ pairs in Au-Au collisions at LHC we scale the $`pp`$ calculations from PYTHIA with the total number of nucleon-nucleon collisions . Typical rapidity densities of 50 $`c\overline{c}`$ pairs were obtained, leading to about 50 $`D`$ and $`\overline{D}`$ mesons and 0.5 primary $`J/\psi `$ at midrapidity. In fig.3 we show the time evolution of the abundance of $`J/\psi `$ mesons as obtained from the solution of the kinetic equations with four different values of the $`J/\psi `$ absorption cross section as described in fig.1. As is seen in fig.3 the yield is very sensitive to the size of the absorption cross section. For example, using the short distance QCD approach leads to negligible secondary production of $`J/\psi `$ mesons. Similar analysis for $`\psi ^,`$ is shown in fig.4. Here, due to the small binding energy of the $`\psi ^,`$ meson and its correspondingly larger size, we have taken the cross section for $`\psi ^,`$ absorption by pions and rho mesons to be energy independent and equal to its geometric value of 10 mb just at threshold . The results in fig.4 show that the secondary yield of $`\psi ^,`$ can be large and reaches the value of almost 1/5 of the initial number of primary $`J/\psi `$. The yield is also seen to be very weakly dependent on the freezeout time. Thermal $`c\overline{c}`$ pairs produced during the evolution of quark-gluon plasma are, within our approach, found to increase the secondary charmonium yield shown in fig<sup>,</sup>s.3,4 by 40$`\%`$ with an equilibrium initial conditions as used in fig.2 and by 20$`\%`$ if using the initial conditions from SSPC for a plasma out of chemical equilibrium. A detailed presentation of the results can be found in . ## 3 Conclusions We have shown that secondary charmonium production in heavy ion collisions appears almost entirely during the mixed phase. The yield of secondarily produced $`J/\psi `$ mesons is very sensitive to the hadronic absorption cross section. Within the context of the short distance QCD approach this leads to negligible values for J/$`\mathrm{\Psi }`$ regeneration. The $`\psi ^,`$ production, however, can be large and may even exceed the initial yield from primary hard scattering. It is thus conceivable that at LHC energy the $`\psi ^,`$ charmonium state can be seen in the final state whereas $`J/\psi `$ production can be entirely suppressed. We acknowledge stimulating discussions with H. Satz and J. Stachel.
no-problem/9908/hep-ph9908219.html
ar5iv
text
# Introduction ### Introduction Hybrid inflation, where some ‘trigger’ field $`\chi `$ responsible for the bulk of the potential is different from the slowly-rolling inflaton field $`\varphi `$, has proved a very useful paradigm which may in the end turn out to be the one chosen by Nature. The original model worked with a tree-level potential, the slope of the inflaton potential being given by the mass term $`\frac{1}{2}m^2\varphi ^2`$. The potential was soon shown to be derivable from spontaneously broken global supersymmetry (susy), with either the $`F`$ term or the $`D`$ term dominating.<sup>2</sup><sup>2</sup>2In these models the inflaton mass vanishes at the level of global supersymmetry, but supergravity corrections can give a suitable mass which indeed tends to be somewhat too large in the $`F`$ term model. For each type of model, the one-loop correction to the inflaton potential coming from the trigger field and its superpartners (trigger supermultiplet) was evaluated . Later, realizations of the tree-level model in the context of softly broken supersymmetry were given , leading to a one-loop correction of a different form . In order to have a more attractive model in the context of supergravity, the dominant loop correction in this case is supposed to come from some gauge supermultiplet, not from the trigger supermultiplet In this note I go beyond specific models. Barring accidental cancellations, the loop correction from the trigger supermultiplet cannot be less than it is in the case of the models with spontaneously broken global supersymmetry. Still barring accidental cancellations, this places a lower bound on the derivative of the inflaton potential, which in turn limits the allowed region of parameter space for a wide class of hybrid inflation models. Assuming that the inflaton fluctuation is the origin of structure in the Universe, we find the bound $$M^4\varphi _{\mathrm{COBE}}\text{ }>\left(10^9\text{GeV}\right)^5,$$ (1) where $`\varphi _{\mathrm{COBE}}`$ is the inflaton field when scales explored by COBE leave the horizon, and $`M\chi `$ is the vacuum expectation value of the trigger field. Whether or not the inflaton fluctuation is the origin of structure, we find the bound $$M^2\varphi _N\text{ }>\left(10^4\text{GeV}\right)^3\sqrt{N},$$ (2) where $`\varphi _N`$ is the inflaton field $`N`$ $`e`$-folds before the end of slow-roll inflation. Implications of these results for TeV-scale inflation will be considered at the end of the paper. Throughout the paper, ‘hybrid inflation’ is taken to mean a model in which the trigger field is fixed during inflation, with the inflaton field moving towards the origin. This excludes ‘inverted’ hybrid inflation in which the inflaton is moving away from origin and ‘mutated’ hybrid inflation in which the trigger field has a time-dependent value adjusted to minimize the potential. Only hybrid inflation in the narrow sense is treated here. ### Basics Let us summarize the basics of inflation model building, as given in say . As usual, $`M_\mathrm{P}(8\pi G)^{1/2}=2.4\times 10^{18}\text{GeV}`$ is the Planck scale while $`H`$ is the Hubble parameter. Our Universe is assumed to be flat so that after it leaves the horizon $`3M_\mathrm{P}^2H^2\rho `$ where $`\rho `$ is the energy density. An overdot denotes differentiation with respect to time and the prime differentiation with respect to the inflaton field $`\varphi `$. We are concerned with the slow-roll paradigm of inflation in which the field equation $`\ddot{\varphi }+3H\dot{\varphi }+V^{}=0`$ is replaced by the slow-roll condition $`3H\dot{\varphi }V^{}`$, and $`V\rho `$ is almost constant on the Hubble timescale. The latter condition requires the flatness condition $$ϵ1,$$ (3) and differentiating the slow-roll condition requires another flatness condition $$|\eta |1,$$ (4) where $`ϵ`$ $``$ $`{\displaystyle \frac{1}{2}}M_\mathrm{P}^2(V^{}/V)^2,`$ (5) $`\eta `$ $``$ $`M_\mathrm{P}^2V^{\prime \prime }/V,.`$ (6) $`N`$ $`e`$-folds before the end of slow-roll inflation, the field value $`\varphi _N`$ is given by $$N=M_\mathrm{P}^2_{\varphi _{\mathrm{end}}}^{\varphi _N}\frac{V}{V^{}}𝑑\varphi ,$$ (7) where $`\varphi _{\mathrm{end}}`$ marks the end of slow-roll inflation. If the inflaton field fluctuation is responsible for structure in the Universe, the COBE measurement of the cosmic microwave background anisotropy requires $$M_\mathrm{P}^3V^{3/2}/V^{}=5.3\times 10^4.$$ (8) This equation applies at the epoch when the distance scale explored by COBE (say $`H_0^1/10`$) leaves the horizon, some number $`N_{\mathrm{COBE}}<60`$ $`e`$-folds before the end of slow-roll inflation. ### Hybrid inflation The original tree-level hybrid inflation model is defined by $$V(\varphi ,\chi )=V_0+\mathrm{\Delta }V(\varphi )\frac{1}{2}|m_\chi ^2|\chi ^2+\frac{1}{2}\lambda ^{}\chi ^2\varphi ^2+\frac{1}{4}\lambda \chi ^4,$$ (9) where $$\mathrm{\Delta }V(\varphi )=\frac{1}{2}m^2\varphi ^2.$$ (10) Inflation takes place in the regime $`\varphi ^2>\varphi _\mathrm{c}^2`$, where $$\varphi _\mathrm{c}|m_\chi |/\sqrt{\lambda ^{}}.$$ (11) In this regime, $`\chi `$ vanishes and the inflaton potential is $$V=V_0+\mathrm{\Delta }V(\varphi ).$$ (12) The constant term $`V_0`$ is assumed to dominate during inflation. The last term of Eq. (9) serves only to determine the vacuum expectation value (vev) of $`\chi `$, achieved when $`\varphi `$ falls below $`\varphi _\mathrm{c}`$. Using that fact that $`V_0`$ vanishes in the vacuum, one learns that the vev is $$\chi M=2V_0^{1/2}/|m_\chi |,$$ (13) and that $$\lambda =\frac{4V_0}{M^4}=\frac{m_\chi ^4}{4V_0}.$$ (14) From Eq. (10), $$\eta =\frac{m^2M_\mathrm{P}^2}{V_0}.$$ (15) It is useful also to define $$\eta _\chi \frac{|m_\chi ^2|M_\mathrm{P}^2}{V_0}=\frac{4M_\mathrm{P}^2}{M^2}.$$ (16) A prompt end to inflation at $`\varphi _\mathrm{c}`$ requires $$\eta _\chi \text{ }>1.$$ (17) As already mentioned, alternative models of hybrid inflation have been proposed where $`\mathrm{\Delta }V`$ is dominated by a loop correction instead of by the mass term Eq. (10). One might also allow significant tree-level terms $`\varphi ^p`$ with $`p>2`$. In any case, once $`\mathrm{\Delta }V`$ is specified, $`\varphi _{\mathrm{COBE}}`$ is determined by Eq. (7) with $$\varphi _{\mathrm{end}}=\mathrm{max}\{\varphi _\mathrm{c},\varphi _{\mathrm{fast}}\},$$ (18) where $`\varphi _{\mathrm{fast}}`$ is the field value when one of Eqs. (3) and (4) fail. Assuming that there are no fine-tuned accidental cancellations, but without specifying the precise form of $`\mathrm{\Delta }V`$, we argue first that the magnitude of the loop correction to $`\mathrm{\Delta }V^{}`$, and hence of $`\mathrm{\Delta }V^{}`$ itself, cannot be much less than $`m_\chi ^4/(16\pi ^2\varphi )`$. Inserting this into Eq. (8) gives the advertised bound Eq. (1), and inserting it into Eq. (7) gives Eq. (2). We go on to show that the same bounds hold for even more general hybrid inflation potentials, and discuss their implication for recent proposals concerning TeV-scale inflation. Further bounds are obtained under additional assumptions. ### The one-loop correction The one-loop correction is $$\mathrm{\Delta }V_{\mathrm{loop}}(\varphi )=\frac{Q^2}{32\pi ^2}\mathrm{Str}^2(\varphi )+\frac{1}{64\pi ^2}\mathrm{Str}\left[^4(\varphi )\left(\mathrm{ln}\frac{^2(\varphi )}{Q^2}\frac{3}{2}\right)\right].$$ (19) Here, $`^2(\varphi )`$ is the field dependent mass-squared matrix for the particles contributing to the loop correction. These particles will in general have spins $`j=0,1/2`$ or $`1`$, and the supertrace is defined as $$\mathrm{Str}A=\underset{j}{}(1)^{2j}(1+2j)\mathrm{Tr}A_j,$$ (20) Here, $`A`$ denotes either $`^2`$ or the square bracket, and $`A_j`$ is the ordinary trace for particles of spin $`j`$. The quantity $`Q`$ is the renormalization scale at which the parameters of the tree-level potential should be evaluated. Its choice is arbitrary, and if all loop corrections were included, the total potential would be independent of $`Q`$ by virtue of the Renormalization Group Equations (RGEs). In any application of quantum field theory, one should choose $`Q`$ so that the total 1-loop correction is small, hopefully justifying the neglect of the multi-loop correction. Unless there is a fine-tuned cancellation the first term of Eq. (19) induces (through the RGEs) a radiative correction to $`m`$ of order $`g_{\mathrm{max}}\mathrm{\Lambda }_{\mathrm{UV}}`$, where $`\mathrm{\Lambda }_{\mathrm{UV}}`$ is the ultra-violet cutoff and $`g_{\mathrm{max}}`$ is a measure of the dominant inflaton coupling. (If this is the coupling to $`\chi `$ and its superpartners, $`g_{\mathrm{max}}=\sqrt{\lambda ^{}}`$.) As is well known, global supersymmetry ensures a precise cancellation, making the first term of Eq. (19) independent of $`\varphi `$ so that it does not contribute to the slope of the potential and does not affect the mass. Supergravity corrections will spoil the cancellation to some extent (see for example ) but since we are interested only in a lower bound on $`|\mathrm{\Delta }V_{\mathrm{loop}}^{}|`$, and are barring accurate accidental cancellations, we ignore such corrections. By the same token we ignore the $`3/2`$ contribution to the second term (which actually depends on the renormalization scheme, the $`3/2`$ holding in the $`\overline{DR}`$ scheme) leaving $$\mathrm{\Delta }V_{\mathrm{loop}}(\varphi )=\frac{1}{64\pi ^2}\mathrm{Str}\left(^4(\varphi )\mathrm{ln}\frac{^2(\varphi )}{Q^2}\right).$$ (21) The contribution of $`\chi `$ is $$\mathrm{\Delta }V_\chi =\frac{1}{64\pi ^2}\left(m_\chi ^4(\varphi )\mathrm{ln}\frac{m_\chi ^2(\varphi )}{Q^2}\right),$$ (22) where $$m_\chi ^2(\varphi )\left(\lambda ^{}\varphi ^2|m_\chi ^2|\right)=\lambda ^{}(\varphi ^2\varphi _\mathrm{c}^2).$$ (23) Making again the optimal assumption of global supersymmetry, the total loop correction from the trigger supermultiplet is $$\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}=\frac{1}{32\pi ^2}\left(m_\chi ^4(\varphi )\mathrm{ln}\left(\frac{m_\chi (\varphi )}{Q}\right)+m_{\stackrel{~}{\chi }}^4(\varphi )\mathrm{ln}\left(\frac{m_{\stackrel{~}{\chi }}(\varphi )}{Q}\right)2m_\mathrm{f}^4(\varphi )\mathrm{ln}\left(\frac{m_\mathrm{f}(\varphi )}{Q}\right)\right),$$ (24) where<sup>3</sup><sup>3</sup>3Eq. (26) corrects Eq. (325) of . $`m_{\stackrel{~}{\chi }}^2(\varphi )`$ $`=`$ $`\lambda ^{}\varphi ^2+m_{\stackrel{~}{\chi }}^2`$ (25) $`m_\mathrm{f}(\varphi )`$ $`=`$ $`\sqrt{\lambda ^{}}\varphi +m_\mathrm{f}.`$ (26) Here $`\stackrel{~}{\chi }`$ denotes the scalar partner, while f denotes the spin-half partner. As will become clear, the minimal value of $`|\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}|`$ is obtained with $`m_\mathrm{f}=0`$ and $`m_{\stackrel{~}{\chi }}^2=|m_\chi ^2|`$. The first condition is usually ensured by some symmetry, and the second will usually be satisfied if global supersymmetry during inflation is broken spontaneously as opposed to softly. Let us consider this case first, and assume for the moment that $`\varphi `$ is significantly bigger than $`\varphi _\mathrm{c}`$. The logs in Eq. (24) can then be taken to have the same argument $`g\varphi /Q`$ and one obtains $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}`$ $`=`$ $`{\displaystyle \frac{m_\chi ^4}{16\pi ^2}}\mathrm{ln}(\sqrt{\lambda ^{}}\varphi /Q)`$ (27) $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}`$ $`=`$ $`{\displaystyle \frac{m_\chi ^4}{16\pi ^2\varphi }}.`$ (28) If susy is broken softly, one will still generally have $`m_\mathrm{f}=0`$ but now $`m_{\stackrel{~}{\chi }}^2`$ can take essentially any value. If it is positive, $`\stackrel{~}{\chi }`$ vanishes both during inflation and in the true vacuum, which as in the previous case justifies its omission from the tree-level potential.<sup>4</sup><sup>4</sup>4To be precise, $`\stackrel{~}{\chi }`$ vanishes during inflation if $`m_{\stackrel{~}{\chi }}V_0^{1/2}/M_\mathrm{P}`$. In the extreme opposite case $`m_{\stackrel{~}{\chi }}^2|m_\chi ^2|`$ the value of $`\stackrel{~}{\chi }`$ is determined by its random quantum fluctuation but it plays no role during inflation. In the intermediate case $`\stackrel{~}{\chi }`$ will be a component of the inflaton field contrary to our assumption that it has only the one component $`\varphi `$. Multicomponent models, as discussed for instance in , are not treated in this note. If $`m_{\stackrel{~}{\chi }}^2`$ is negative, the omission of $`\stackrel{~}{\chi }`$ from the tree-level potential is strictly justified only if $`|m_{\stackrel{~}{\chi }}|=|m_\chi |`$, its effect then being the trivial replacement $`\chi ^2\chi ^2+\stackrel{~}{\chi }^2`$. However, its omission is justified in practice more generally. Indeed, if $`|m_{\stackrel{~}{\chi }}|<|m_\chi |`$, the field $`\stackrel{~}{\chi }`$ is held at the origin until the field $`\chi `$ is destabilized, making the former ineffective except for a modest increase in the value of $`V_0`$.<sup>5</sup><sup>5</sup>5Again we discount very small values of $`|m_{\stackrel{~}{\chi }}|`$, which would lead to a different model of inflation. The opposite case may be discounted because it is equivalent to an interchange of the labels $`\chi `$ and $`\chi ^{}`$. For simplicity we assume $`m_{\stackrel{~}{\chi }}^2=m_\chi ^2`$, noting that $`|\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}|`$ will be at least as big in other cases. Still assuming that $`\varphi `$ is appreciably bigger than $`\varphi _\mathrm{c}`$ we find in the softly broken case $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}`$ $`=`$ $`{\displaystyle \frac{\lambda ^{}}{8\pi ^2}}m_\chi ^2\varphi ^2\mathrm{ln}(\sqrt{\lambda ^{}}\varphi /Q)`$ (29) $`|\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}|`$ $`=`$ $`\left|{\displaystyle \frac{\lambda ^{}}{8\pi ^2}}m_\chi ^2\varphi \left(2\mathrm{ln}(\sqrt{\lambda ^{}}\varphi /Q)+1\right)\right|\left|{\displaystyle \frac{\lambda ^{}}{8\pi ^2}}m_\chi ^2\varphi \right|.`$ (30) The final equality is valid if $`Q`$ is chosen to make $`\mathrm{ln}(\sqrt{\lambda ^{}}\varphi /Q)`$ (and therefore $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}`$) vanish for the value of $`\varphi `$ under consideration. Any choice making $`|\mathrm{ln}(\sqrt{\lambda ^{}}\varphi /Q)|1`$ would be equally valid in that it would still justify the neglect of higher loop corrections, and one could actually choose $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}=0`$ at (say) the COBE scale. But as already remarked, the total potential (tree level plus loop correction) is independent of $`Q`$ by virtue of the RGE’s. We shall now argue that the right hand side of Eq. (30) provides a lower limit on the slope of the total potential. The one-loop correction will be typically be valid in an interval over which $`\varphi `$ varies by a factor of a few, if $`Q`$ is set equal to some $`\sqrt{\lambda ^{}}\varphi `$ in the range. At the one-loop level, $`\stackrel{~}{\mathrm{\Delta }V}_{\chi \stackrel{~}{\chi }\mathrm{f}}\frac{1}{2}m^2(Q)\varphi ^2+\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}(Q,\varphi )`$ is approximately independent of $`Q`$ by virtue of the one-loop RGE for $`m^2`$. One can write $`\stackrel{~}{\mathrm{\Delta }V}_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}=(\lambda ^{}/4\pi ^2)m_\chi ^2\varphi \mathrm{ln}(\varphi /\varphi _0)`$, where $`\varphi _0`$ is some constant. In general this makes $`\stackrel{~}{\mathrm{\Delta }V}_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}`$ of the advertised magnitude Eq. (30). If $`\varphi _0`$ is within the range of validity of this expression, $`\stackrel{~}{\mathrm{\Delta }V}_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}`$ vanishes at $`\varphi =\varphi _0`$ and will be reduced by a factor $`|1\varphi /\varphi _0|`$ if $`\varphi `$ is very close to $`\varphi _0`$. Let us see what happens in that case. For hybrid inflation $`\varphi _0`$ should correspond to a maximum, with $`\varphi _{\mathrm{end}}<\varphi <\varphi _0`$. From Eq. (7) we learn that $`(1\varphi _{\mathrm{end}}/\varphi _0)=e^{|\eta |N}(1\varphi _N/\varphi _0)`$, where $`\eta M_\mathrm{P}^2V^{\prime \prime }/V`$ is evaluated at $`\varphi _0`$. (We assume that the expression for $`\stackrel{~}{\mathrm{\Delta }V}_{\chi \stackrel{~}{\chi }\mathrm{f}}`$ is valid down to $`\varphi _{\mathrm{end}}`$ which is good enough for a crude order of magnitude estimate .) We are considering the case that $`\varphi _N/\varphi _0=1`$ with extreme accuracy; but then, the flatness condition $`|\eta |1`$ requires that in the physically relevant range $`N\text{ }<N_{\mathrm{COBE}}<60`$, $`\varphi _{\mathrm{end}}/\varphi _0`$ is also equal to 1 with extreme accuracy. As always, we assume that such accidental fine-tuning does not occur. The conclusion is that indeed the right hand side of Eq. (30) will provide a lower bound on the slope of $`\stackrel{~}{\mathrm{\Delta }V}_{\chi \stackrel{~}{\chi }\mathrm{f}}`$. Finally, consider the case that $`\varphi `$ is very close to $`\varphi _\mathrm{c}`$. The loop contribution from $`\chi `$ is very small, but those of its superpartners are unsuppressed. Focussing say on the contribution of $`\stackrel{~}{\chi }`$, it is easy to check that with (say) $`Q`$ chosen to make $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}=0`$ one finds a a result for $`|\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}|`$ at least as big as Eq. (28).<sup>6</sup><sup>6</sup>6To be precise, the slope of the renormalization-group-improved tree-level potential will be at least as big as the right hand side of Eq. (28). As in the previous case, one could actually choose $`Q`$ to make $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}`$ vanish at a given value of $`\varphi `$. ### Constraints on hybrid inflation models Discounting as always the possibility of a precise accidental cancellation, one will have $$|\mathrm{\Delta }V^{}|\text{ }>|\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}|.$$ (31) Using the minimal estimate Eq. (28), the COBE normalization Eq. (8) with Eq. (16) then gives $$M^4\varphi _{\mathrm{COBE}}\text{ }>5.4\times 10^5M_\mathrm{P}^3V_0^{1/2}=\left(10^9\text{GeV}\right)^5\left(\frac{V_0}{1\text{MeV}}\right)^{\frac{1}{2}}.$$ (32) To obtain the advertised final inequality Eq. (1), we set $`V_0^{1/4}\text{ }>1\text{MeV}`$, the smallest possible value since reheating must occur before nucleosynthesis. In the case of soft susy breaking Eq. (30) provides a stronger lower bound on $`\mathrm{\Delta }V^{}`$, which with Eq. (8) gives $$5\times 10^4M_\mathrm{P}^3\text{ }<\frac{V_0^{3/2}}{\lambda ^{}|m_\chi ^2|\varphi }.$$ (33) Using $`\varphi _{\mathrm{COBE}}>\varphi _\mathrm{c}`$ gives a bound on $`M`$ alone, $$M^3\text{ }>5.4\times 10^5\sqrt{\lambda ^{}}M_\mathrm{P}^3.$$ (34) This bound is less interesting than Eq. (32) because it is evaded by the case of spontaneous symmetry breaking, and because it can be made arbitrarily weak by lowering $`\lambda ^{}`$. Although it may not be very natural, there does not seem to be any bar to taking $`\lambda ^{}`$ very small. In particular, reheating could take place through the coupling of the inflaton to some field other than the trigger field. We can obtain further results if the original model $`\mathrm{\Delta }V\frac{1}{2}m^2\varphi ^2`$ is supposed to be valid. In this case, $$\varphi _{\mathrm{COBE}}^2=e^{2\eta N}\varphi _\mathrm{c}^2\varphi _\mathrm{c}^2.$$ (35) (The approximate equality corresponds to $`e^{2\eta N}1`$ which is good enough for order-of-magnitude estimates.) The COBE normalization Eq. (8) becomes $`\lambda ^{}`$ $`=`$ $`2.8\times 10^7e^{2\eta N}\eta ^2\eta _\chi `$ (36) $``$ $`3\times 10^7\eta ^2\eta _\chi .`$ (37) For this original model to be valid one needs $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{}m^2\varphi `$ and $`\mathrm{\Delta }V_{\chi \stackrel{~}{\chi }\mathrm{f}}^{\prime \prime }m^2`$, for all $`\varphi _\mathrm{c}\text{ }<\varphi \text{ }<\varphi _{\mathrm{COBE}}`$. These two constraints are about the same, and may be written $$\frac{\lambda ^{}}{32\pi ^2}\frac{\eta _\chi }{\eta }1.$$ (38) We can use the COBE normalization Eq. (37) to eliminate any one of the three parameters. Eliminating $`\eta `$ reproduces Eq. (34). Eliminating $`\eta _\chi `$, Eq. (38) becomes $$\lambda ^{}(\eta /22)^{3/2}.$$ (39) This is weaker than Eq. (34) if $`M/M_\mathrm{P}\text{ }<0.005(\eta /.025)^{1/4}`$. Finally, eliminating $`\lambda ^{}`$ gives $$\eta (90M/M_\mathrm{P})^4.$$ (40) In all of this we considered the epoch of inflation that is supposed to generate large scale structure, through the inflaton field perturbation. One can also consider an epoch of late slow-roll inflation, lasting only a few $`e`$-folds, whose only purpose it to dilute unwanted relics. In that case the bounds we have considered disappear, but a significant bound still comes from the requirement of $`N\text{ }>1`$ $`e`$-folds of inflation. Inserting Eqs. (31) and (27) into Eq. (7) gives indeed $$\varphi ^2\text{ }>\frac{2\sqrt{\lambda }V_0^{1/2}}{\lambda ^{}}+\frac{\lambda NM_\mathrm{P}^2}{2\pi ^2}.$$ (41) The first term is just $`\varphi _\mathrm{c}^2`$, and is typically negligible. Combining this equation with Eq. (14) gives $$M^2\varphi _N\text{ }>\left(10^4\text{GeV}\right)^3\left(\frac{V_0^{1/4}}{1\text{MeV}}\right)^2\sqrt{N}.$$ (42) Since $`V_0^{1/4}>1\text{MeV}`$, this leads to the advertised bound Eq. (2). ### Nonrenormalizable terms The original tree-level potential Eq. (9) ignores nonrenormalizable terms. They are of the form $`\lambda _{mn}\mathrm{\Lambda }_{\mathrm{UV}}^{4mn}\varphi ^m\chi ^n`$, with $`m+n5`$, where $`\mathrm{\Lambda }_{\mathrm{UV}}`$ is the ultra-violet cutoff for the effective field theory relevant during inflation. These terms summarize the physics which is ignored by the effective field theory. Since quantum gravity certainly becomes significant on Planck scales one must have $`\mathrm{\Lambda }_{\mathrm{UV}}\text{ }<M_\mathrm{P}`$, but $`\mathrm{\Lambda }_{\mathrm{UV}}`$ will be smaller if the effective field theory breaks before Planck scales are reached. This could happen in three ways. First, a different field theory may take over, containing fields that have been integrated out in the effective theory. Second, the scale of quantum gravity could be lower than $`M_\mathrm{P}`$ because there are extra dimensions with a large compactification radius. Third, in the presence of a large compactification radius field theory may give way to string theory well below the scale of quantum gravity. The last two possibilities have received a lot of attention lately, with the focus on the lowest conceivable cutoff $`\mathrm{\Lambda }_{\mathrm{UV}}\text{TeV}`$. The coefficients $`\lambda _{mn}`$ are in principle determined by the theory that takes over on scales bigger than the cutoff. For generic fields they will be roughly of order 1, at least for moderate values of $`n`$ and $`m`$, without any cancellation between different non-renormalizable terms.<sup>7</sup><sup>7</sup>7For very large $`n`$ and/or $`m`$ one might expect factors like $`1/n!`$ . In the context of string theory it is known that certain fields, such as the dilaton, exist for which this is not the case. Although one cannot rule out the possibility that some of these fields might be suitable for inflation, attempts to use them for that purpose have not so far been successful when one imposes the COBE normalization .<sup>8</sup><sup>8</sup>8One of these fields has been proposed as the trigger field in a $`D`$-term inflation model but the required properties have not been derived from the relevant string theory. Assuming that the non-renormalizable terms indeed have couplings of order 1, both the inflaton field and the trigger field will have to be $`\text{ }<\mathrm{\Lambda }_{\mathrm{UV}}`$ in magnitude if these terms are to be under control.<sup>9</sup><sup>9</sup>9We differ here from Kanti and Olive , who in considering chaotic inflation with large extra dimensions require only $`\varphi \text{ }<M_\mathrm{P}`$. If Eq. (32) holds, this is possible only if $$\mathrm{\Lambda }_{\mathrm{UV}}\text{ }>10^9\text{GeV}\left(V_0^{1/4}/1\text{MeV}\right)^{2/5}.$$ (43) In any case one has from Eq. (42), $$\mathrm{\Lambda }_{\mathrm{UV}}\text{ }>10^4\text{GeV}\left(V_0^{1/4}/1\text{MeV}\right)^{2/3}.$$ (44) Even with the relevant field values below $`\mathrm{\Lambda }_{\mathrm{UV}}`$, one or more non-renormalizable terms could be significant. In particular, either or both of the interaction terms in Eq. (9) can be replaced by a non-renormalizable term . Replacing them both gives $$V(\varphi ,\chi )=V_0+\mathrm{\Delta }V(\varphi )\frac{1}{2}|m_\chi ^2|\chi ^2+\frac{1}{2}\lambda _n^{}\mathrm{\Lambda }_{\mathrm{UV}}^{2n}\varphi ^n\chi ^2+\frac{1}{4}\lambda _m\mathrm{\Lambda }_{\mathrm{UV}}^{4m}\chi ^m,$$ (45) with $`n>2`$ and $`m>4`$. The original model is recovered for $`n=2`$ and $`m=4`$. Let us see what difference these replacements make. Replacing $`\frac{1}{4}\lambda \chi ^4`$ is trivial, because Eq. (14) still holds approximately which means that Eqs. (32)–(40) and Eq. (42) remain valid. The replacement of $`\frac{1}{2}\lambda ^{}\varphi ^2\chi ^2`$ by $`\frac{1}{2}\lambda _n^{}\mathrm{\Lambda }_{\mathrm{UV}}^{2n}\varphi ^n\chi ^2`$ is less trivial. Eq. (11) is replaced by $$\lambda _n^{}\mathrm{\Lambda }_{\mathrm{UV}}^{2n}\varphi _\mathrm{c}^n=|m_\chi ^2|.$$ (46) In Eq. (24) one has $`m_\chi ^2(\varphi )=\lambda _n^{}\mathrm{\Lambda }_{\mathrm{UV}}^{2n}\varphi ^n+m_\chi ^2`$ and similarly for $`m_{\stackrel{~}{\chi }}^2(\varphi )`$ and $`m_\mathrm{f}(\varphi )`$. Discounting factors of order $`n`$, the main result Eq. (32) remains valid, and so does Eq. (42), while Eq. (34) becomes $$\frac{M^3}{M_\mathrm{P}^3}\text{ }>5.4\times 10^5\left(\frac{m_\chi }{\mathrm{\Lambda }_{\mathrm{UV}}}\right)^{1\frac{2}{n}}\left(\lambda _n^{}\right)^{\frac{1}{n}}.$$ (47) Finally, Eq. (39) becomes $$(\lambda _n^{})^{\frac{2}{n}}\left(\frac{m_\chi }{\mathrm{\Lambda }_{\mathrm{UV}}}\right)^{2\frac{4}{n}}(\eta /22)^{3/2}.,$$ (48) while Eq. (40) is unchanged. Requiring $`M\text{ }<\mathrm{\Lambda }_{\mathrm{UV}}`$ in Eq. (47) (with $`\lambda _n^{}1`$) leads to a stronger condition than Eq. (43).<sup>10</sup><sup>10</sup>10In advocating the condition $`M\mathrm{\Lambda }_{\mathrm{UV}}`$ we differ from Randall et al. , who advocate $`MM_\mathrm{P}`$ even in the presence of the ultra-violet cutoff. With $`n=4`$ one has $$\mathrm{\Lambda }_{\mathrm{UV}}\text{ }>10^{12}\text{GeV}.$$ (49) Finally we mention a different generalization of the original model , which is to introduce an interaction $`A\varphi \chi ^2`$ which dominates the term $`\frac{1}{2}|m_\chi ^2|\chi ^2`$ . This amounts to the replacement $`|m_\chi ^2|A^2/\lambda ^{}`$, which does not affect the main result Eq. (32). ### Inflation at the TeV scale Although hybrid inflation remains at least as promising as the alternative of single-field inflation, there are significant constraints on the parameter space. As an important example, we consider recent proposals for hybrid inflation at and below the TeV scale ($`V_0^{1/4}1\text{TeV}`$). In an era of TeV-scale inflation is invoked to provide a new model of baryogenesis, in which the Higgs field is created by the oscillation of the inflaton field. To achieve this, the trigger field is identified with the electroweak Higgs field which requires $`\lambda 1`$ in Eq. (9), corresponding to $`M\text{TeV}`$. The loop correction is ignored to produce an apparently viable tree-level model with $`\varphi _{\mathrm{COBE}}\text{TeV}`$, satisfying the COBE normalization. But Eq. (32) shows that with the loop correction included, any COBE-normalized model in which the Higgs is the trigger field will have $`\varphi _{\mathrm{COBE}}\text{ }>\varphi _\mathrm{c}\text{ }>10^{33}\text{GeV}`$, a value presumably too large to contemplate. Even if we abandon the COBE normalization by assuming that the TeV-scale inflation lasts only a few $`e`$-folds, Eq. (41) requires $`\varphi _{\mathrm{COBE}}M_\mathrm{P}`$. With such a value non-renormalizable terms are likely to spoil inflation, and even if they were ignored one would have to see if this completely different model of baryogenesis is still viable. Alternatively, one might explore the possibility that the Higgs couples to the inflaton in a non-hybrid model, still taking on board the fact that the coupling to the Higgs will induce a loop correction to the inflaton potential. To summarize briefly, the specific model of baryogenesis written down in is invalidated by the loop correction, and it remains to be seen whether alternatives in the same spirit can be constructed. The model we just looked at assumes that quantum gravity is at the usual scale $`M_\mathrm{P}`$. The rest of this note deals with the possibility that it is much lower, focussing on the lowest conceivable scale $`\mathrm{\Lambda }_{\mathrm{UV}}\text{TeV}`$. Kaloper and Linde explore the possibility of COBE-normalized hybrid inflation in this case. They invoke the COBE-normalized tree-level model with $`\lambda \lambda ^{}1`$, corresponding to $`m_\chi M\varphi _{\mathrm{COBE}}V_0^{1/4}\text{TeV}`$ and (because of the COBE normalization) $`m\text{ }<10^7H`$. They point out that the low value of $`m`$ makes the model rather unattractive, because supergravity will typically generate $`m`$ at least of order $`H`$. We have seen that with the above parameter choice the loop correction will invalidate the tree-level model, which means that the above difficulty does not in fact exist; the bound $`m\text{ }<10^7H`$ does not in fact apply. However, Eq. (32) points to another problem; any COBE-normalized model of hybrid inflation with $`\mathrm{\Lambda }_{\mathrm{UV}}\text{TeV}`$ is problematic because when the loop correction is included the inflaton and/or the trigger field have to be at least of order $`10^{10}\text{GeV}`$. The bottom line here is that we concur with the conclusion of Kaloper and Linde, that hybrid inflation with quantum gravity at the TeV scale looks very problematic. As mentioned earlier, Halyo takes the view that field values $`\text{TeV}`$ are capable of being justified, by making the trigger field a superstring modulus. He advocates the usual $`D`$-term hybrid inflation model ($`\mathrm{\Delta }V`$ given by Eq. (27)), with the quantum gravity scale of order TeV, and with a very small coupling $`\lambda 10^{27}`$ derived from a superstring theory with large extra dimensions. In apparent contradiction with our general bound Eq. (32), he concludes that an ‘initial’ inflaton field value (defined as the value when $`N=4\times 10^5`$) and $`M`$ can both be of order $`10^6\text{GeV}`$. This conflict can be traced to the fact that Halyo uses a wrong formula for the COBE normalization. On the reasonable assumption $`\varphi _{\mathrm{COBE}}\varphi _\mathrm{c}`$, the correct formula is $$V^{1/4}=\left(\frac{50}{NC}\right)^{1/4}\left(\frac{\lambda }{4}\right)^{1/4}\times 6\times 10^{15}\text{GeV},$$ (50) where $`C`$ is the number of charged pairs coupling to the inflaton.<sup>11</sup><sup>11</sup>11Note that the $`g`$ in Eq. (246) of should actually be $`g^{1/2}`$. However, this formula confirms the basic tenet of ; that with a very small $`\lambda `$ the $`D`$-term hybrid inflation model can give inflation at the TeV scale, if one accepts field values much bigger than this scale. Let us end this discussion by noting that potentials of the ‘new’ inflation, as opposed to hybrid inflation, type can give COBE-normalized TeV-scale inflation with the inflaton field $`\text{ }<\text{TeV}`$. Such potentials are of the form $`V=V_0+\mathrm{\Delta }V`$, with $`V_0`$ dominating and $`\mathrm{\Delta }V<0`$. It is known that COBE-normalized TeV-scale inflation can be achieved for the cases $`\mathrm{\Delta }V\lambda \varphi ^4\mathrm{ln}(\varphi /\varphi _0)`$ (with $`\varphi \varphi _0`$) , $`\mathrm{\Delta }V=c\mathrm{\Lambda }_{\mathrm{UV}}^2\varphi ^6+b\mathrm{\Lambda }_{\mathrm{UV}}^4\varphi ^8`$ and $`\mathrm{\Delta }V=\lambda \varphi ^4+b\mathrm{\Lambda }_{\mathrm{UV}}^{4q}\varphi ^q`$ . The middle case invokes no small couplings $`(cb1)`$. The first and third cases require $`\lambda 10^{15}`$; however, as we have seen the first case has to come (in the context of global supersymmetry) from a non-renormalizable interaction like $`\lambda _6\mathrm{\Lambda }_{\mathrm{UV}}^2\varphi ^6\chi ^2`$, leading after soft supersymmetry breaking to $`\lambda =(\lambda _6/8\pi ^2)(m_\chi /\mathrm{\Lambda }_{\mathrm{UV}})^2`$, and one could have $`\lambda _61`$. For any kind of slow-roll inflation, the inflaton mass during inflation has to be much less than $`H=V_0^{1/2}/M_\mathrm{P}`$ in order to satisfy the flatness condition $`\eta 1`$ (barring a fine-tuned accidental cancellation between the mass term and the displayed terms). After inflation it becomes much bigger, both in the above models and in hybrid inflation, so that there need be no problem with reheating. ### Further remarks on inflation with quantum gravity at the TeV scale All of this assumes that the size of the extra dimensions, responsible for quantum gravity at the TeV scale, have their present size while cosmological scales leave the horizon during inflation. It has been shown that by the end of inflation, they must have their present size, with an accuracy $`10^{14}(T_{\mathrm{RH}}/10\text{MeV})^{3/2}`$ where $`T_{\mathrm{RH}}`$ is the reheat temperature. Otherwise, the moduli corresponding to the extra dimensions would be overproduced afterwards. To achieve this accuracy one generally needs $`V^{1/4}\text{TeV}`$ by the end of inflation.<sup>12</sup><sup>12</sup>12During inflation the canonically normalized modulus $`\chi `$, corresponding to the size of the extra dimensions, is displaced from its minimum by $`\delta \chi =\sqrt{n/(n+2)}2H^2M_\mathrm{P}/m_\chi ^2`$, where $`m_\chi `$ is its mass and $`n`$ is the number of extra dimensions. This agrees with a rough earlier estimate $`\delta \chi H^2M_\mathrm{P}/m_\chi ^2`$, made for moduli in general. (The upper bound on $`V^{1/4}`$ depends on the number of extra dimensions and the moduli mass.) While cosmological scales are leaving the horizon during inflation, the extra dimensions must be stabilized, since significant variation would spoil the observed scale independence of the spectrum of the primordial curvature perturbation. The simplest hypothesis is that they remain stabilized thereafter, so that they indeed have their present value while cosmological scales leave the horizon. An alternative is to assume that the extra dimensions are stabilized, while cosmological scales are leaving the horizon, with sizes much smaller than at present. In that case, after going to the four-dimensional Einstein frame, the scale of quantum gravity will be far above the TeV scale and it becomes easier to construct a COBE-normalized model of inflation. This hypothesis of ‘asymmetric’ inflation has the disadvantage of invoking two separate epochs of stabilization for the extra dimensions. (It should be carefully distinguished from the idea that the extra dimensions must be very small at the beginning of inflation, long before cosmological scales leave the horizon. This indeed seems to be desirable on quite general grounds .) Irrespective of the asymmetric inflation hypothesis, or indeed of the scale of quantum gravity, one may need a few $`e`$-folds of late inflation to get rid of unwanted relics (including moduli) produced after the slow-roll inflation that generates structure. Suppose first that the late inflation is hybrid inflation. The COBE constraint Eq. (32) on a hybrid inflation model disappears, but Eq. (42) remains. Assuming that all field values should be $`\text{ }<\mathrm{\Lambda }_{\mathrm{UV}}`$, this requires $`\mathrm{\Lambda }_{\mathrm{UV}}\text{ }<10^4N^{1/3}\text{GeV}`$. A hybrid inflation model saturating this bound can provide a few $`e`$-folds of inflation with quantum gravity at the scale $`\mathrm{\Lambda }_{\mathrm{UV}}10\text{TeV}`$. It is clear that a model saturating the bound would have $`\mathrm{\Delta }V`$ given by Eq. (27), corresponding to spontaneously broken global susy with supergravity corrections assumed to be negligible. The usual $`D`$-term model is of this kind, and indeed Halyo has already noted the fact that a few $`e`$-folds of such inflation is possible, with both quantum gravity and the field values at the $`10\text{TeV}`$ scale. One should however remember that even in the $`D`$-term (as opposed to an $`F`$-term ) model, non-renormalizable terms in the superpotential may be a problem if $`\varphi `$ is at the scale $`\mathrm{\Lambda }_{\mathrm{UV}}`$ . What one really needs to be comfortable is $`\varphi \mathrm{\Lambda }_{\mathrm{UV}}`$, which would require $`\mathrm{\Lambda }_{\mathrm{UV}}10^4\text{GeV}`$. Instead of hybrid inflation one may consider a ‘new inflation’ potential or some other slow-roll potential . A completely different alternative is thermal inflation . In this paradigm, the zero-temperature inflaton potential is $`V_0\frac{1}{2}m^2\varphi ^2`$, with $`|m|V_0^{1/4}`$, but unsuppressed couplings lead to a thermal correction of order $`T^4+T^2\varphi ^2`$. In the regime $`m\text{ }<T\text{ }<V_0^{1/4}`$, some $`N\mathrm{ln}(V_0^{1/4}/m)`$ $`e`$-folds of thermal inflation can occur, with $`\varphi `$ held at the origin, after which $`\varphi `$ reverts to its vev $$\varphi V_0^{1/2}/m.$$ (51) The expected decay time of the inflaton is $`\mathrm{\Gamma }^110^2\varphi ^2/m^3`$. This corresponds to a reheat temperature $$T_{\mathrm{RH}}\left(\frac{\text{TeV}}{\varphi }\right)^{\frac{5}{2}}\left(\frac{V_0^{\frac{1}{4}}}{100\text{MeV}}\right)^3\times 3\text{MeV}.$$ (52) This can satisfy the cosmological constraints, including the nucleosynthesis requirement $`T_{\mathrm{RH}}>\text{MeV}`$. We see that thermal inflation, originally proposed in the context of quantum gravity at the usual scale $`M_\mathrm{P}`$, may be viable also in the context of TeV-scale gravity.<sup>13</sup><sup>13</sup>13It has been invoked in this context by Dvali in the special case that the inflaton corresponds to the separation between $`D`$-branes, but the above discussion following the original papers shows the required properties of the inflaton are model-independent, and in no way special. The main difference from the usual case is that the required mass $`m`$ has no obvious explanation. (In the usual case one invokes a mass $`100\text{GeV}`$, which is natural in the context of gravity-mediated supersymmetry breaking.) We end by revisiting an earlier paper , which concluded that TeV-scale quantum gravity does not avoid the need for supersymmetry. This conclusion was based on the fact that the inflaton mass during slow-roll inflation has to be much less than $`HV_0^{1/2}/M_\mathrm{P}V_0^{1/4}\text{ }<\mathrm{\Lambda }_{\mathrm{UV}}`$. We would like to point that some apparent escape routes from that conclusion can now be closed off. The first of these is the possibility that the extra dimensions during inflation might be much smaller than at present . As we seen, explicit calculation has shown that this cannot be the case during, at least, the last few $`e`$-folds. The second is that the last few $`e`$-folds of inflation might be thermal as opposed to slow-roll. As we have seen, this too would require an inflaton mass orders of magnitude below the TeV scale. Another possibility might be to have slow-roll inflation, with all couplings of the inflaton very small. Then the non-supersymmetric radiative correction $`\delta mg_{\mathrm{max}}\mathrm{\Lambda }_{\mathrm{UV}}`$ might be small enough without any need of supersymmetry. In that case , reheating could take place only by the gravitational production of particles, with mass much less than $`HV_0^{1/2}/M_\mathrm{P}`$ and unsuppressed couplings.<sup>14</sup><sup>14</sup>14We discount here the possibility that the couplings of the inflaton could somehow be suppressed during inflation and unsuppressed afterwards. There does not seem to be any known way of achieving this. Such particles would require supersymmetry to protect their masses. Barring unforseen escape routes, it seems that even with quantum gravity at the TeV scale, one needs during (at least) the last few $`e`$-folds of inflation a very light scalar particle, whose mass will have to be protected by supersymmetry. This means , contrary to what was initially hoped , that there is no particular reason why Nature should have placed the scale of quantum gravity at the TeV scale, as opposed to somewhere else in the range $`\text{TeV}\text{ }<\mathrm{\Lambda }_{\mathrm{UV}}\text{ }<M_\mathrm{P}`$. Indeed, while the TeV choice removes the need for supersymmetry in the Higgs sector of the theory , the need reappears in the inflaton sector. The conclusion is that we are unlikely to observe quantum gravity at a future collider, though of course we may be lucky! ## Acknowledgements I am indebted to Toni Riotto, Andrei Linde and Andre’ Lukas for useful discussions.
no-problem/9908/math-ph9908011.html
ar5iv
text
# Instantaneous Reflection and Transmission Coefficients and a Special Method to Solve Wave Equation ## I. INTRODUCTION In the cases of quantum mechanical barrier problem reflection and transmission coefficients can be calculated. In those cases with the interaction of the potential field, one fraction of incident particle (or wave) is reflected back to the infinity (reflection coefficient) and other fraction transmits into other side or may tunnel through the barrier (transmission coefficient). If the barrier is of constant height and depth, the reflection and transmission coefficient can be easily calculated . In the case of varying potential, from the far away of potential field, reflection and transmission coefficient can be calculated using WKB approximation . In this present paper we will show how the reflection and transmission of the particle changes in the cases of varying potential point to point. Due to variation of potential, in each point particle should feel different potential field so the fraction of particle (or wave) which reflects and transmits will be changing at each point. Here we will calculate the analytical space dependent expression of reflection and transmission coefficients which indicates transmission and reflection of the particle at any arbitrary point with respect to the transmission of the particle at the immediate previous point. So the local values of incident and reflected wave amplitude can be found. With this, we shall present the solution of the differential equation (i.e. Schrödinger equation), containing that space dependent reflected and transmitted amplitude. Secondly, in our daily problems of physics and mathematics there are several differential equations which can not be solved analytically. If we want to solve those equations we need numerical methods. If, however, the nature of the coefficients of derivative terms or inhomogeneous term namely factor, due to presence of those factors corresponding differential equation to be unsolvable, are well behaved then using the methods which will be explained here the differential equation can be solved analytically using WKB approximation method. Actual reason of non-analyticity is that corresponding factors can not be integrated analytically. We can consider, a particle is moving in a potential field whose analytical form is complicated. Obviously corresponding Schrödinger equation will be non-analytical. Although the reflection and transmission coefficients for that particle can be calculated but the wave function can not be analytically calculated exactly, because for this job we need to integrate analytically the wave vector $`k`$, which includes the complicated potential. In next section we will discuss how the space dependent reflection and transmission coefficient can be calculated for the case of varying potential. In §3, we will establish approximate analytical solution of a Schrödinger equation with the presence of potential with complicated analytical form. In §4 we will illustrate an example where our method can be applied for finding solution and reflection-transmission coefficients. In §5 we will make our conclusions. ## II. Method to find local reflection and transmission coefficients We consider an one dimensional Schrödinger like differential equation as $$\frac{d^2y}{dx^2}+k^2(x)y=0$$ $`(1)`$ where, $`k^2(x)=EV(x)`$, $`E=`$ total energy, $`V(x)=`$ potential energy. (2) Here, the independent variable $`x`$ is varying from $`\mathrm{}`$ to $`\mathrm{}`$, i.e., nothing but cartesian co-ordinate. Using WKB method the solution of this equation is $$y(x)=\frac{A}{\sqrt{k}}exp(iu)+\frac{B}{\sqrt{k}}exp(iu)$$ $`(3)`$ where, $$u=k(x)𝑑x,$$ $`(4)`$ $`A`$ and $`B`$ are constants of integration. The reflection and transmission coefficients calculated by WKB method means the coefficients are calculated at a certain point, that means from a certain point what fraction of particle is transmitted inside and what other fraction is reflected outside. If we calculate this transmission and reflection of the particle from different points, results should be different. So the probability of transmission and reflection of the particle from point to point will be different for the case of varying potential. So the constants $`A`$ and $`B`$ should be changed from point to point due to change of effective potential which is felt by the particle. Here we will give an analytical space dependent expression for transmission and reflection coefficient by which one can find, how the reflection and transmission probability change from point to point for the case of varying potential. It is assumed that far away from the potential field (at $`x=\mathrm{}`$), the potential barrier height is almost constant, so the reflection and transmission coefficients are almost same. In those regions we safely can choose $`A`$ and $`B`$ as pure constants. Again, all along the sum of transmission and reflection coefficients should be unity, so far away from the potential field the relations between $`A`$ and $`B`$ are, $$A=B+c(c=\mathrm{constant}),$$ $`(5)`$ $$A^2+B^2=k.$$ $`(6)`$ It is very easy to check from incident and reflected current that transmission and reflection coefficients are $`\frac{A^2}{k}`$ and $`\frac{B^2}{k}`$ respectively. Solving these two equation we get, $$A(x)=\frac{c}{2}+\frac{\sqrt{2kc^2}}{2}$$ $`(7a)`$ $$B(x)=\frac{c}{2}+\frac{\sqrt{2kc^2}}{2}$$ $`(7b)`$ These $`A(x)`$ and $`B(x)`$ are very slowly varying functions at large $`x`$. These are indicating, at far away from the potential barrier (where potential is varying in very slow manner), how constants $`A`$ and $`B`$ are slowly changing. So from these slowly varying $`A`$ and $`B`$, one can find out how transmission and reflection coefficients are varying at far distance. Now, constant $`c`$ can be calculated from the boundary conditions. From pure WKB approximation constant $`c`$ can be calculated. From WKB method, far away from the potential field reflection and transmission coefficients and corresponding reflected and transmitted amplitude can be calculated. So constant $`c`$ is immediately determined. Obviously, at $`x=\mathrm{}`$ these expressions of $`A`$ and $`B`$ will not be valid because of simplified assumptions. At the minimum value of $`x`$, i.e., $`x=\mathrm{}`$, we have to put another condition. It is clear that the potential field, which is felt by the particle, extended in between $`\mathrm{}`$ to $`\mathrm{}`$ and assumed that at $`x=\mathrm{}`$ barrier height goes down to zero or constant value i.e., the potential is asymptotically flat. So the particle which reaches at $`x=\mathrm{}`$, should feel free. As a result, there transmission should be $`100\%`$ and corresponding reflection is zero (hereafter, this will be called as inner boundary condition). By introducing this inner boundary condition in eqn. (7) we get, $$C(x)=c_1+A(x)=c_1+\frac{c}{2}+\frac{\sqrt{2kc^2}}{2},$$ $`(8a)`$ $$D(x)=c_2+B(x)=c_2\frac{c}{2}+\frac{\sqrt{2kc^2}}{2}.$$ $`(8b)`$ Here, constants $`c_1`$ and $`c_2`$ are introduced to modify the reflection and transmission coefficients according to the inner boundary condition. Here, one necessary condition is the sum of reflection and transmission coefficients should be one. By modifying the coefficients, it is seen that $$C^2(x)+D^2(x)=\left(c_1+\frac{c}{2}\right)^2+\left(c_2\frac{c}{2}\right)^2+(c_1+c_2)\sqrt{2kc^2}+\frac{(2kc^2)}{2}=h(x)[say]=\left[\frac{h(x)}{k(x)}\right]k(x).$$ $`(9)`$ So, it is advisable to choose the modified coefficients of the wave function as follows: $$a(x)=\frac{C(x)}{\sqrt{h/k}}$$ $`(10a)`$ $$b(x)=\frac{D(x)}{\sqrt{h/k}}$$ $`(10b)`$ so that $$a^2(x)+b^2(x)=k(x)$$ $`(11)`$ and the transmission and reflection coefficients are $`\frac{a^2(x)}{k(x)}`$ and $`\frac{b^2(x)}{k(x)}`$ respectively which are explicitly written as: $$T(x)=\frac{(c_1+\frac{c}{2})}{h(x)}\left(c_1+\frac{c}{2}+\sqrt{2k(x)c^2}\right)+\frac{2k(x)c^2}{4h(x)}$$ $`(12a)`$ $$R(x)=\frac{(c_2\frac{c}{2})}{h(x)}\left(c_2\frac{c}{2}+\sqrt{2k(x)c^2}\right)+\frac{2k(x)c^2}{4h(x)}.$$ $`(12b)`$ Now we describe, how to determine the constants $`c_1`$ and $`c_2`$. First of all to satisfy $`B(x=\mathrm{})=0`$ (inner boundary condition) $`c_2`$ is determined and fixed. Then introducing this $`c_2`$ in $`h(x)`$ and equating the expression of modified reflection coefficient \[$`b^2(x)/k(x)`$\] and reflection coefficient valid only at $`x=\mathrm{}`$ \[$`B^2(x)/k(x)`$, which is correct at far away from the black hole\] we get the other constant $`c_1`$. These expressions for transmission and reflection coefficients are valid in whole region. The space dependent coefficients of incident and reflected waves, $`a(x)`$ and $`b(x)`$ are valid for arbitrary $`x`$. So the final form of the solution using WKB method is $$y(x)=\frac{a(x)}{\sqrt{k}}exp(iu)+\frac{b(x)}{\sqrt{k}}exp(iu).$$ $`(13)`$ Here one important thing is to be noted that, WKB approximation is valid only if $`\frac{1}{k}\frac{dk}{dx}<<k`$. So one can calculate the space dependent reflection and transmission coefficient only if potential vary in such a manner that $`k`$ satisfies the above condition. Otherwise our method to find out space dependent amplitudes and coefficients is not valid because our method is based on WKB approximation. Equation (13) is valid at any point. It is indicating the solution and corresponding reflection and transmission coefficients for an arbitrary $`x`$ as if at each point WKB approximation method is applied instantaneously. So our modified WKB method can be called as Instantaneous WKB Method or IWKB method . For the cases, where $`\frac{1}{k}\frac{dk}{dx}<<k`$ is not satisfied, the potential can be replaced by a large number of square steps and space dependent reflection and transmission coefficients can be found . Here, corresponding problem is reduced to as simple quantum mechanical barrier problem with multiple steps where at each step junction the wave functions and its derivatives of two separate regions should be continuous. Detail discussion is in Mukhopadhyay & Chakrabarti . ## III. Solution of the equation In the previous section, although we have given the form of the solution with space dependent incident and reflected coefficients but the Eiconal $`u`$ still yet to be determined. If the analytical form of the potential and corresponding wave vector $`k`$ is well integrable analytically then $`u`$ can be calculated immediately but sometimes the form of $`k`$ may be complicated such that it can not be integrated analytically then $`u`$ can not be evaluated unless very special cases are chosen. In this section we want to give an approximate solution of eqn. (1) using WKB (actually IWKB, as mentioned in last section) approximation even if the analytical form of $`k`$ is not well integrable. As we mentioned in INTRODUCTION that if $`k^2(x)`$ is well behaved then we can give an analytical form of the eqn. (1) using the following trick. If the nature of the coefficients of derivative and zeroth order derivative terms are well behaved but analytically complicated then those analytical form of the coefficients can be replaced by piecewise continuous analytical function with simple form and of same behaviour as exact complicated functions. After this replacement $`k`$ will be of simple form and we can integrate it to find out $`u`$. Once we know the analytical form of $`u`$, we can present the analytical solution of the equation. In different ways we can replace the complicated functions. Any function can be expanded in terms of polynomial. It is seen that if the corresponding polynomial is of order two then the corresponding $`k`$ can be integrated easily. But using one polynomial of order two, in the total range, the replacement of the complicated function can not be done. So in different regions, different coefficients of the polynomial are chosen i.e., different second order polynomials are chosen such that at the boundaries (where the polynomials change), value of the two different polynomials and its derivatives are same. In this way, by using many number of piecewise polynomials the analytically complicated well behaved functions can be replaced. If in the differential eqn. (1), the analytical form of $`V`$ and corresponding $`k`$ is of non-analytical form (complicated analytical expression) then potential $`V`$ can be replaced by piecewise continuous polynomial as $$V_l(x)=a_l+b_lx+c_lx^2$$ $`(14)`$ and the corresponding $`k(x)`$ is defined as $$k_l^2(x)=(Ea_l)b_lxc_lx^2.$$ $`(15)`$ Integrating this we can find out analytical form of $`u`$ as $$u_l(x)=2c_lk_lc_l\sqrt{(Ea_l)}log\left|\frac{k_l\sqrt{(Ea_l)}}{k_l+\sqrt{(Ea_l)}}\right|+\mathrm{constant}$$ $`(16)`$ Where index $`l`$ in the coefficients of the polynomial is indicating that in different ranges of $`x`$ different values of $`a`$, $`b`$, $`c`$ may be chosen. This type of replacement also can be done by some other well integrable functions according to the nature of the complicated functions. In the next section we will give an example, where, to solve the differential equation with the potential of complicated analytical form, we replace the complicated function (potential) by the function of simple analytical form otherthan second order polynomials. In this way we can give the analytical form of the solution of Schrödinger equation like equation (3) even if the analytical form of the potential is complicated. Since we have used the WKB approximation method (more clearly IWKB method), our solution only will be valid for $`\frac{1}{k}\frac{dk}{dx}<<k`$. If the potential vary in such a way that above condition is not being satisfied then the solution of the Schrödinger or Schrödinger like equation is not possible by this method. For those cases we have to employ some other methods, one of them is indicated in last section (by reducing the problem to simple quantum mechanical barrier problem). Sometimes for those cases, the Schrödinger equation can be reduced into Bessel equation and solution is possible by expressing in terms of Airy functions . ## IV. illustrative example Dirac equation in Kerr geometry can be separated into radial and angular parts \[4-5\]. Then corresponding decoupled radial equation on some transformation of independent variable can be reduced into Schrödinger like equation as $$\frac{d^2Z}{dx^2}+\left(\sigma ^2V(x)\right)=0.$$ $`(17)`$ In comparison with Schrödinger equation with $`\text{}h=c=G=1`$, we can say $`V(x)`$ is nothing but potential and $`\sigma ^2`$ is proportional to energy of the particle. Here the analytical form of the potential is complicated as $$V=\frac{\mathrm{\Delta }^{\frac{1}{2}}(\text{}\lambda ^2+m_p^2r^2)^{3/2}}{[\omega ^2(\text{}\lambda ^2+m_p^2r^2)+\text{}\lambda m_p\mathrm{\Delta }/2\sigma ]^2}[\mathrm{\Delta }^{\frac{1}{2}}(\text{}\lambda ^2+m_p^2r^2)^{3/2}+((rM)(\text{}\lambda ^2+m_p^2r^2)+3m_p^2r\mathrm{\Delta })]$$ $$\frac{\mathrm{\Delta }^{\frac{3}{2}}(\text{}\lambda ^2+m_p^2r^2)^{5/2}}{[\omega ^2(\text{}\lambda ^2+m_p^2r^2)+\text{}\lambda m_p\mathrm{\Delta }/2\sigma ]^3}[2r(\text{}\lambda ^2+m_p^2r^2)+2m_p^2\omega ^2r+\text{}\lambda m_p(rM)/\sigma ],$$ $`(18)`$ where $`a=`$ Kerr parameter, $`m_p=`$ mass of the incident particle, $`\sigma =`$ frequency of the incident particle, $`M=`$ mass of the black hole, $`\text{}\lambda =`$ separation constant, $`r=`$ radial co-ordinate, $`\mathrm{\Delta }=r^22Mr+a^2`$, $`\omega =r^2+a^2+am/\sigma `$, $`x=f(r)`$. For details see . Now, although the analytical form of the potential is complicated but for particular set of physical parameter (such as, $`a,\sigma ,m_p`$ etc.) the nature of $`V`$ is well behaved. So we can replace it by suitable peace-wise continuous analytical function of simple form. We can choose it as second order polynomial as explained in previous section. But for convenience we choose here in different form. To show the nature of the potential, we choose as an example one set of physical parameter, such as: $`a=0.5`$; $`M=1`$; $`m_p=0.8`$; orbital quantum number, $`l=\frac{1}{2}`$; azimuthal quantum number, $`m=\frac{1}{2}`$; $`\sigma =0.8`$; $`\text{}\lambda =0.92`$ from . The simple analytical form which is chosen to replace the complicated form of the potential for this set of physical parameter is given as $$V_l(x)=a_l+b_lexp\left(\frac{x}{c_l}\right)$$ $`(19)`$ and corresponding $`k`$ and eiconal $`u`$ are given as $$k_l(x)=\sqrt{(\sigma ^2a_l)b_lexp(\frac{x}{c_l})},$$ $`(20)`$ $$u(x)=2c_lk_l(x)+c_l\sqrt{(\sigma ^2a_l)}log\left|\frac{\sqrt{(\sigma ^2a_l)}+k_l(x)}{\sqrt{(\sigma ^2a_l)}k_l(x)}\right|+\mathrm{constant}.$$ $`(21)`$ The behaviour of the functions of eqn. (18) and (19) are same. In case of (19) in different ranges of $`x`$ different values of $`a,b,c`$ are chosen. For these physical parameters, coefficients $`a,b,c`$ of mapping function in different ranges of $`x`$ are given as: $`a=0`$, $`b=0.187354`$, $`c=3.75`$ for $`x`$ $`\mathrm{}`$ to $`0`$, $`a=0.603`$, $`b=0.415646`$, $`c=8.79`$ for $`x`$ $`0`$ to $`30`$, $`a=0.629`$, $`b=0.12690038`$, $`c=26.3`$ for $`x`$ $`30`$ to $`109`$, $`a=0.63543098`$, $`b=0.037193439`$, $`c=73.5`$ for $`x`$ $`109`$ to $`208`$, $`a=0.63543098`$, $`b=0.2228925`$, $`c=45`$ for $`x`$ $`208`$ to $`310`$, etc. So finally we can say the analytical form of the solution of the Schrödinger equation is possible for this set of physical parameter. Similarly for other sets of physical parameter one can find out solution following same method (by mapping the complicated analytical form of the potential by the simple form which may be polynomial or analytical function like eqn. (19) and which are analytically integrable). The space dependent transmission and reflection coefficients can be calculated following the method explained earlier. Now we will calculate the constants $`c,c_1,c_2`$ by imposing boundary conditions for the potential of given set of physical parameter. Using general WKB approximation method, from the far away of the black hole (say at $`x=310`$) the reflection and transmission coefficients of the particle for the potential shown in Fig. 1 can be calculated as $`T=0.299`$ and $`R=0.701`$. Using these values of the coefficient constant $`c`$ can be calculated as $$c=\sqrt{Tk}\sqrt{Rk}=0.0913$$ $`(22)`$ where $`k`$ is the value of wave number at $`x=310`$. Using this $`c`$ when we calculate the transmitted and reflected amplitude from (7a-b) (which are valid at large distance from the black hole) obviously it will not satisfy the inner boundary condition. To satisfy the inner boundary condition first $`c_2`$ is introduced to reduce $`B(x)`$ to zero at $`x=\mathrm{}`$ as $`c_2=0.6765`$. Then as explained in section II, by matching the reflection amplitudes of eqn. (7b) and (10b) at infinity i.e., $$b(x)=B(x),$$ $`(23)`$ $`c_1`$ can be evaluated as $`c_1=0.1639`$. Now all the constants are known. In Fig. 2 variation of instantaneous reflection and transmission coefficients are shown. It is seen that as the potential barrier lower down the transmission probability increases and reflection probability decreases. For the cases of other set of physical parameters where the potential barrier may be of different type, using the same method constants can be calculated. In the final solution with the insertion of the value of constants it can be checked that WKB approximation is still valid except in the regions where $`\sigma ^2`$ is close to $`V`$ and this modified WKB (IWKB) solution satisfies the original differential equation if $`\frac{1}{k}\frac{dk}{dx}<<k`$. ## V. conclusions Here, we show how reflection and transmission coefficients vary for the particle moving in spatially varying potential field. Actually we solve the Quantum Mechanical barrier problem, where the barrier height changes point to point. For solution, we introduce modified WKB approximation method namely instantaneous WKB method i.e., IWKB method where at each point, instantaneously, the WKB approximation method can be applied and corresponding reflection and transmission coefficients can be calculated. Since the potential changes point to point corresponding reflection and transmission coefficients change. We also indicate how Schrödinger equation is solved analytically where coefficients of derivative terms are of complicated analytical form. It is shown that if coefficients are well behaved then the the analytical solution can be set up by suitably mapping the coefficient-functions to functions of simple form. In this way, any such kind of differential equation can have analytical form of the solution which could be useful for further study. ## VI. acknowledgment It is a great pleasure to the author to thank Prof. Sandip K. Chakrabarti for many serious discussion regarding this work and support to write this paper. ## Figure Caption Fig. 1 : Variation of potential $`V`$ (solid curve) with respect to radial distance $`x`$ and total energy $`E`$ (dashed curve) of particle. Fig. 2 : Variation of transmission (solid curve) and reflection coefficients (dashed curve) with respect to radial distance $`x`$.
no-problem/9908/nucl-th9908025.html
ar5iv
text
# Nuclear Physics on the Light Front ## 1 INTRODUCTION Much of nuclear physics is concerned with the transition between nucleon-meson and quark-gluon degrees of freedom, with deep-inelastic lepton-nucleus scattering, Drell-Yan production and high energy (e,e’p) and (p,pp) reactions being some of the tools of investigation. Since high energies are involved, it is absolutely necessary to incorporate relativity. Our project is to implement Poincare invariance in nuclear physics by using light front dynamics. This is useful in analyzing experiments because the canonical momentum variable, $`k^+`$, is closely related to experimental observables. The challenge is to use these variables while incorporating the dynamical richness necessary to properly describe nuclei. Consider the plus-momentum distributions $`f_{B,M}(k^+)`$ which give the probability that a nuclear baryon $`B`$ or meson $`M`$ has a plus-momentum of $`k^+`$. If the relevant nuclear wave functions depend on $`k^+`$, the canonical spatial variable is $`x^{}=x^0x^3`$. This leaves $`x^+=x^0+x^3`$ to be used as a time variable. Thus the light front dynamics is very different than the usual dynamics. If this light front formalism is used, $`f_{B,M}`$ are simply related to the square of the ground state nuclear wave function. Thus the use of light front dynamics provides vast simplicities, provided one can obtain the ground state wave function using realistic Lagrangians and many-body techniques. It is easier to carry out such a program for the effective hadronic Lagrangians of nuclear physics than for $`QCD`$ because hadrons all have a non-zero mass and because the vacuum is not a condensed state of nucleon-anti-nucleon pairs. Another advantage is that the Fock state expansion of the wave function is in terms of on-shell particles. The remainder is meant to present a brief outline of the progress. The main goal is to provide a series of examples to show that the light front can be used for high energy realistic-nuclear physics. ## 2 LIGHT FRONT QUANTIZATION OF HADRONIC LAGRANGIANS You have to have a Lagrangian $``$ no matter how bad. We use ones in which the degrees of freedom are nucleons, vector and scalar mesons and pions. The existence of $``$ allows the derivation of the canonical-symmetric energy momentum tensor $`T^{\mu \nu }`$. In light front dynamics the momentum is $`P^+=P^0+P^3`$ where $`P^\mu `$ is the total momentum operator: $`P^\mu =\frac{1}{2}d^2x_{}𝑑x^{}T^{+\mu }`$. The $`x^+`$ development operator is $`P^{}`$. One necessary detail is that that $`T^+`$ must be expressed in terms of independent degrees of freedom. One uses the equations of motion to express the dependent degrees of freedom in terms of the independent ones, and uses these constraint equations in the expression for $`T^+`$. Except for chiral Lagrangians, the necessary quantization procedure has been carried out long ago by other authors-. ## 3 INFINITE NUCLEAR MATTER IN MEAN FIELD APPROXIMATION: PLUS-MOMENTUM DISTRIBUTIONS The first problem we tried to solve is that of understanding nuclear matter in the mean field approximation using the Walecka model. This problem has been solved in a manifestly covariant manner, so that the existing results for the energy of the system provide a useful check of our light front procedure. We are able to reproduce the standard results . The interesting feature of light front dynamics that one is able to calculate the plus-momentum distributions. For this simple problem we found that the nuclear vector mesons carry about a third of the nuclear plus momentum, but their momentum distribution has support only at $`k^+=0`$, with $`k^+f_v(k^+)=0.35M_N\delta (k^+)`$. Thus the vector mesons do not contribute to nuclear deep inelastic scattering. This zero mode effect occurs because in the mean field approximation, the meson fields of an infinite system are constant in both space and time. This is an intriguing result which is caused by the large values of the scalar and vector potentials, which are characteristic of the Walecka model. The nucleons carry only 65% of the nuclear plus-momentum, a result in severe disagreement with deep inelastic scattering data. It is necessary to see if it survives for the case of finite-sized nuclei, and if correlations between nucleons are included. ## 4 FINITE NUCLEI IN MEAN FIELD APPROXIMATION We showed that the necessary variational principle is a constrained one which fixes the expectation value of the total momentum operator $`P^+`$ to be the same as that for $`P^{}`$. This is the same as minimizing the sum of the total momentum operators: $`P^{}+P^+`$. A new light-front version of the equation for the single nucleon modes was obtained, and a new numerical technique for their solution was introduced. The ground state wave function is treated as a meson-nucleon Fock state, and the meson fields are treated as expectation values of field operators in that ground state. The resulting equations for these expectation values was shown to be closely related to the usual meson field equations. The computed binding energies are essentially the same as for the usual equal-time theory. The nucleon plus momentum distribution $`f_N(k^+)`$ peaks for $`k^+`$ about seventy percent of the nucleon mass, which again is far too small to be consistent with deep inelastic scattering data. The mesonic component of the ground state wave function was used to determine the scalar and vector meson momentum distribution functions, and the vector mesons were found to carry about thirty percent of the nuclear plus-momentum. We are currently investigating other Lagrangians which lead to smaller magnitudes of scalar and vector potentials and therefore yield better descriptions of the deep inelastic scattering data. ## 5 PIONS AND CHIRAL SYMMETRY Light front quantization of a chiral Lagrangian can be performed , if one uses a Lagrangian due to Gürsey. Pion-nucleon scattering at tree level was shown to reproduce soft pion theorems ## 6 NUCLEON-NUCLEON SCATTERING The Weinberg-type equation is the light front version of the Lippman-Schwinger equation. This equation is equivalent to the the Blankenbecler-Sugar equation, except that retardation effects need to be included. Our hadronic chiral Lagrangian was used to obtain a light front version of a one-boson-exchange nucleon-nucleon potential (OBEP). The accuracy of our description of the nucleon-nucleon (NN) data is good, and similar to that of other relativistic OBEP models. ## 7 NUCLEAR MATTER WITH NUCLEON-NUCLEON CORRELATIONS The trivial nature of the vacuum in the light front formalism was exploited in deriving the equations analogous to the the Hartree-Fock and Brueckner Hartree-Fock equations. Applying our light front OBEP, the nuclear matter saturation properties are reasonably well reproduced. The computed value of the compressibility, 180 MeV, is smaller than that of alternative relativistic approaches to nuclear matter in which the compressibility usually comes out too large. We showed that replacing the meson degrees of freedom by a NN interaction is a consistent approximation, and that the formalism allows one to calculate corrections to this approximation in a well-organized manner. The mesonic Fock space components of the nuclear wave function are studied also, and aspects of the meson and nucleon plus-momentum distribution functions are computed. We find that there are about 0.05 excess pions per nucleon. The magnitudes of the scalar and vector potentials are far smaller than found in the mean field approximation. If we neglect the influence of two-particle-two-hole states to approximate $`f(k^+)`$ the nucleons are found to carry 81% (as opposed to the 65% of mean field theory) of the nuclear plus momentum. This represents a vast improvement in the description of nuclear deep inelastic scattering as the minimum value of the ratio $`F_{2A}/F_{2N}`$ is increased by a factor of twenty towards the data, which is not enough to provide a satisfactory description. We can be optimistic about future results because including nucleons with momentum greater than $`k_F`$ would substantially increase the computed ratio $`F_{2A}/F_{2N}`$ since $`F_{2N}(x)`$ decreases very rapidly with increasing values of $`x`$ and because $`M^{}`$ would increase at high momenta Turn now to the experimental information about the nuclear pionic content. The Drell-Yan experiment on nuclear targets showed no enhancement of nuclear pions within an error of about 5%-10% for their heaviest target. Understanding this result is an important challenge to the understanding of nuclear dynamics . Here we have a good description of nuclear dynamics, and our 5% enhancement is consistent, within errors, with the Drell-Yan data. ## 8 SUMMARY The light front approach has now been applied to infinite nuclear matter in the mean field approximation, finite-sized nuclei in the same approximation, $`\pi N`$ and $`NN`$ scattering, and to correlated nucleons in infinite nuclear matter Thus it seems that one can use the light front approach to compute nuclear energies, wave functions and the experimentally important plus-momentum distributions. There are indications that the computed quantities will ultimately be in good agreement with experiment. But the use of light front dynamics in nuclear physics is only in its infancy, and much remains to be done to understand nuclear deep inelastic data the future expected flood of data on the (e,e’p) and (p,pp) reactions.
no-problem/9908/hep-th9908104.html
ar5iv
text
# References EDO-EP-28 August, 1999 Mass Hierarchy from Many Domain Walls Ichiro Oda <sup>1</sup><sup>1</sup>1 E-mail address: ioda@edogawa-u.ac.jp. Edogawa University, 474 Komaki, Nagareyama City, Chiba 270-0198, JAPAN ## Abstract We construct a new model with exponential mass hierarchy by starting with the Einstein-Hilbert action with the cosmological constant in five dimensions plus an action describing many domain walls in four dimensions. The model includes many hidden sectors and one visible sector, and each four-dimensional domain wall, that is, 3-brane, interacts with one another through only a gravitational interaction and realizes many universe cosmology inspired by D-brane perspective. It is shown that in the present model only even numbers of domain walls are allowed to locate in five dimensional space-time and the validity of Randall-Sundrum scenario, which explains mass hierarchy between the Planck mass and the electro-weak scale in our world, depends on a relative relation between our world and hidden worlds. Recently, there has been considerable interest and activity in the higher dimensional theories to solve the hierarchy problem. It is well known that in the Standard Model based on the gauge group $`SU(3)\times SU(2)\times U(1)`$ there exist two scales, those are, the electro-weak scale $`m_{EW}10^3`$ GeV and the Planck scale $`M_{Pl}10^{19}`$ GeV. The hierarchy problem is a problem concerning the two mass scales. The light Higgs mass $`m_{EW}10^3`$ GeV is needed in the Standard Model, but the quadratic divergences at the loop levels and the renormalization effects associated with the light Higgs field cause this electro-weak mass to have the very huge Planck mass $`M_{Pl}10^{19}`$ GeV, which corresponds to the cut-off scale of the ultra-violet divergences. Thus, solving the hierarchy problem is equivalent to explaining the vast disparity and quantum stability of the two scales without any fine tuning of parameters at each perturbation level. In recent works, it has been pointed out that the existence of large extra compact spatial dimensions might give us a solution for the hierarchy problem . The key idea is the following: in the simple case, the four-dimensional Planck scale $`M_{Pl}`$ and the $`(4+n)`$-dimensional Planck scale $`M_{}`$ have a relation $`M_{Pl}^2=M_{}^{n+2}R_c^n`$ where $`R_c`$ is a compactification radius. Even if $`M_{}`$ is around the order of the electro-weak scale, $`M_{Pl}`$ can become the Planck scale by taking $`R_c`$ to be large enough. But against this naive expectation, it was found that in this scenario the mass hierarchy is merely converted into another hierarchy problem between the electro-weak scale and the compactification radius. Subsequently, a new scenario has been proposed for the hierarchy problem without appealing to large extra compactification dimensions . According to this new scenario, the large hierarchy of the Planck scale and the electro-weak scale is explained by the nontrivial redshift factor arisinging from the boundary condition where an extra dimension has a structure of an orbifold $`S^1/Z_2`$. The model used in the analysis is obtained by M-theory compactified on an $`S^1/Z_2`$ and a deformed Calabi-Yau threefold and has been examined in detail by several groups . The above model describes our world as one of four-dimensional two slabs in five-dimensional space-time while the other slab is regarded as a hidden world which interacts with our world by a gravitational interaction. Although the model is quite of interest from the perspective of M-theory as mentioned above, it might be possible to construct a new model based on the idea of D-branes where an arbitrary number of paralell D-branes are located at various places in the fifth dimensional direction in the bulk and interact with one another by a gravitational force. Indeed, such a model also appears to be welcome from cosmology since in modern cosmology our universe (visible 3-brane) is conjectured to be not a unique and selected universe but have been created with other many universes (hidden 3-branes) simultaneously on an equal footing in the big bang era. Thus it is interesting to construct a new model with one visible world and many hidden worlds and ask how mass hierarchy (and cosmological constant) problem is resolved. The purpose of this paper is to pursue this line of thoughts, present such a concrete model and then show that the validity of a scenario proposed by Randall-Sundrum explaining mass hierarchy depends on a relative relation between our world and hidden worlds. As a bonus, it turns out that one disadvantage of the previous model, the necessity for a singular orbifold geometry , will disappear in the present model since the existence of solutions of the Einstein equation demands us to choose a smooth manifold $`S^1`$ in place of a singular orbifold geometry $`S^1/Z_2`$ when there are more than two branes in a theory. We start with the Einstein-Hilbert action with the cosmological constant in five dimensions plus an action describing $`n`$ domain walls in four dimensions : $`S={\displaystyle \frac{1}{2\kappa ^2}}{\displaystyle d^4x_0^{2L}𝑑z\sqrt{g}\left(R+2\mathrm{\Lambda }\right)}+{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle _{z=L_i}}d^4x\sqrt{g_i}_i,`$ (1) where the cosmological constant $`\mathrm{\Lambda }`$ is taken to a positive number, which implies that the geometry of five-dimensional bulk is anti-de Sitter space-time. (Note that we have changed the sign in front of the cosmological constant compared to a conventional choice so the positive cosmological constant corresponds to the anti-de Sitter space-time.) The fifth dimension $`z`$ is compact with the size $`2L`$. Moreover, $`\kappa `$ denotes the five-dimensional gravitational constant with a relation $`\kappa ^2=8\pi G_N=\frac{8\pi }{M_{}^2}`$ where $`G_N`$ and $`M_{}`$ are the five-dimensional Newton constant and the five-dimensional Planck scale, respectively. In this article we follow the standard conventions of the textbook of Misner, Thorne and Wheeler . It is edifying to mention two differences between our starting action and that of previous works . An obvious difference is that we have introduced not two domain walls but $`n`$ many domain walls located at $`L_i`$ in the $`z`$ direction.<sup>2</sup><sup>2</sup>2In recent work , many branes are also considered from a different context. (Precisely speaking, all $`z=L_i`$ are not always occupied by domain walls. As shown later, the Einstein equation fixes locations of domain walls completely.) Another important difference lies in the fact that in the previous M-theory model the two domain walls are located at the boundaries which are the fixed points of an orbifold $`S^1/Z_2`$, whereas in the present D-brane model the fifth dimension has a topology of $`S^1`$ so that various domain walls are not located at the boundary. Variation of the action (1) with respect to the five-dimensional metric tensor leads to the Einstein equation: $`\sqrt{g}\left(R^{MN}{\displaystyle \frac{1}{2}}g^{MN}R\right)=\sqrt{g}g^{MN}\mathrm{\Lambda }+\kappa ^2{\displaystyle \underset{i=1}{\overset{n}{}}}\sqrt{g_i}g_i^{\mu \nu }\delta _\mu ^M\delta _\nu ^N_i\delta (zL_i),`$ (2) where $`M,N,\mathrm{}`$ denote five-dimensional indices, whereas $`\mu ,\nu ,\mathrm{}`$ do four-dimensional ones. In deriving this equation, we have neglected the contributions from the Lagrangians $`_i`$ of domain walls. In other words, the domain wall actions play a role as sources only providing the cosmological constant in the bulk. In order to solve the equation (2) in an analytical way, we require the following metric ansatz in such a way that the red-shift factor and expanding universe are taken into account: $`ds^2`$ $`=`$ $`g_{MN}dx^Mdx^N`$ (3) $`=`$ $`u(z)^2\left(dt^2+v(t)^2d\stackrel{}{x}^2\right)+b(z)^2dz^2,`$ where $`\stackrel{}{x}`$ denotes the three-dimensional spatial coordinates. Under the ansatz (3), the Einstein equation (2) reduces to three combined differential equations: $`6\left({\displaystyle \frac{u^{}}{u}}\right)^23\left({\displaystyle \frac{b}{u}}\right)^2\left[{\displaystyle \frac{\ddot{v}}{v}}+\left({\displaystyle \frac{\dot{v}}{v}}\right)^2\right]=b^2\mathrm{\Lambda },`$ (4) $`3\left[{\displaystyle \frac{u^{\prime \prime }}{u}}+\left({\displaystyle \frac{u^{}}{u}}\right)^2\right]3{\displaystyle \frac{b^{}}{b}}{\displaystyle \frac{u^{}}{u}}3\left({\displaystyle \frac{b}{u}}\right)^2\left({\displaystyle \frac{\dot{v}}{v}}\right)^2=b^2\mathrm{\Lambda }+\kappa ^2b{\displaystyle \underset{i=1}{\overset{n}{}}}_i\delta (zL_i),`$ (5) $`3\left[{\displaystyle \frac{u^{\prime \prime }}{u}}+\left({\displaystyle \frac{u^{}}{u}}\right)^2\right]3{\displaystyle \frac{b^{}}{b}}{\displaystyle \frac{u^{}}{u}}\left({\displaystyle \frac{b}{u}}\right)^2\left[2{\displaystyle \frac{\ddot{v}}{v}}+\left({\displaystyle \frac{\dot{v}}{v}}\right)^2\right]=b^2\mathrm{\Lambda }+\kappa ^2b{\displaystyle \underset{i=1}{\overset{n}{}}}_i\delta (zL_i),`$ (6) where the prime and the dot denote a differentiation with respect to $`z`$ and $`t`$, respectively. Now from Eqs. (5), (6), it is straightforward to derive the equation for variable $`v`$ $`{\displaystyle \frac{\ddot{v}}{v}}=\left({\displaystyle \frac{\dot{v}}{v}}\right)^2,`$ (7) which can be integrated to $`v(t)=v(0)e^{Ht},`$ (8) where $`H`$ denotes the Hubble constant of expanding universe . The implications of the present model to cosmology will be examined in a separate paper so we shall take $`v(t)=1`$ from now on in order to focus our attention on mass hierarchy. Moreover, we set $`b(t)=r_c`$ where $`r_c`$ is a constant. Under the simplified ansatz $`v(t)=1`$ and $`b(t)=r_c`$, Eqs.(4), (5) (or (6)) take rather simple forms $`\left({\displaystyle \frac{u^{}}{u}}\right)^2=k^2r_c^2,`$ (9) $`{\displaystyle \frac{u^{\prime \prime }}{u}}=k^2r_c^2+{\displaystyle \frac{\kappa ^2}{3}}r_c{\displaystyle \underset{i=1}{\overset{n}{}}}_i\delta (zL_i),`$ (10) where we have defined $`k=\sqrt{{\displaystyle \frac{\mathrm{\Lambda }}{6}}}.`$ (11) Now we wish to find a solution which satisfies the above Einstein equations (9), (10). However, it seems to be difficult to find a general solution, so instead we aim to find some special solutions closely relating to a model with exponential mass hierarchy, whose concrete expression is given by $`u(z)=e^{kr_cf(z)},`$ (12) where $`f(z)`$ is a certain function of variable $`z`$ to be determined in what follows. The most effective and easiest way of determining the function $`f(z)`$ is to construct it in a direct manner according to the following procedure. First, divide the region $`0<z<2L`$ into $`(n1)`$ small pieces $`L_i<z<L_{i+1}(i=1,2,\mathrm{},n1)`$. Second, combine two adjacent pieces into one pair in order from $`L_1=0`$ or $`L_n=2L`$, from which we have $`\frac{n1}{2}`$ distinct pairs. (We assume that $`n`$ is an odd number for a while. This issue will be argued later.) Third, draw a piecewisely continuous straight line with a slope $`+1`$ or $`1`$ in each pair by turns. Finally, since the topology of the fifth dimension is a circle $`S^1`$ the two boundaries $`z=L_10`$ and $`z=L_n2L`$ are identified by setting the periodic boundary condition. Then except one point mentioned shortly, it is straightforward to show that the line obtained in this way satisfies the Einstein equations (9), (10) by using the Fourier series expansion. Here one would like to discuss an important subtlety associated with the definition of $`\epsilon (0)`$. In showing that Eq.(12) with $`f(z)`$ obtained in the above procedure satisfies the Einstein equations (9), (10) one encounters the step function $`\epsilon (x)`$ defined as $`\epsilon (x)={\displaystyle \frac{|x|}{x}}=\{\begin{array}{cc}+1,\hfill & \text{for }x>0\hfill \\ 1,\hfill & \text{for }x<0\hfill \end{array}.`$ (13) At some $`L_i`$ on which domain walls sit, we have an ambiguous quantity $`\epsilon (0)`$, for which we usually define as $`\epsilon (0)=0`$ or $`\epsilon (0)=1`$. The validity of the Einstein equation now requires us to choose $`\epsilon (0)=1`$. Thus at first sight we do not meet any inconsistency, but various results should not depend on such an ambiguous quantity. Indeed, in the situation at hand this quantity has a close connection with a physical fact, namely, the existence of singularities at the location of domain walls, so a careful treatment is needed. In this article, in order to avoid an ambiguous quantity $`\epsilon (0)`$, we wish to regularize our model in a such way that we introduce an infinitely small thickness $`\mathrm{\Delta }`$ of domain walls and take a limit $`\mathrm{\Delta }0`$ after all calculations. Under this regularization, any ambiguity never occur. Let us present two interesting solutions. These solutions describe even domain walls standing along $`S^1`$ at some intervals in five dimensional space-time. One solution is given by $`f(z)`$ $`=`$ $`|z|+{\displaystyle \underset{i=1}{\overset{\frac{n1}{2}}{}}}(1)^i|zL_{2i}|L,`$ $`f^{}(z)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{\frac{n1}{2}}{}}}(1)^i\epsilon (zL_{2i})+1,`$ $`f^{\prime \prime }(z)`$ $`=`$ $`2{\displaystyle \underset{i=1}{\overset{\frac{n1}{2}}{}}}(1)^i\delta (zL_{2i}),`$ (14) where $`L_i`$ satisfies the relations $`L_{2i}={\displaystyle \frac{L_{2i1}+L_{2i+1}}{2}},L_1=0,L_n=2L,`$ (15) with $`i=1,2,\mathrm{},\frac{n1}{2}`$. Moreover, $`_i(i=1,2,\mathrm{},\frac{n1}{2})`$ must satisfy the relations $`_{2i}=(1)^{i+1}{\displaystyle \frac{6k}{\kappa ^2r_c}}.`$ (16) This solution expresses $`\frac{n1}{2}`$ even domain walls locating at $`L_{2i}`$. In this paper, we would like to take account of one visible 3-brane and many hidden 3-branes more than or equal to 1 so $`n`$ runs over $`5,9,13,\mathrm{}`$. Another interesting solution takes the form $`f(z)`$ $`=`$ $`{\displaystyle \underset{i=2}{\overset{n1}{}}}(1)^{i+1}|zL_i|+L,`$ $`f^{}(z)`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{n1}{}}}(1)^{i+1}\epsilon (zL_i)1,`$ $`f^{\prime \prime }(z)`$ $`=`$ $`2{\displaystyle \underset{i=1}{\overset{n1}{}}}(1)^{i+1}\delta (zL_i),`$ (17) where $`L_i`$ also satisfies Eq.(15). This time, $`_i(i=1,2,\mathrm{},\frac{n1}{2})`$ must satisfy the relations $`_1`$ $`=`$ $`_n={\displaystyle \frac{3k}{\kappa ^2r_c}},`$ $`_{2i}`$ $`=`$ $`_{2j+1}={\displaystyle \frac{6k}{\kappa ^2r_c}},`$ (18) where $`0<j<i`$. Now the number of domain walls is an even number $`n1`$ ($`n`$ takes values $`3,5,7,\mathrm{}`$), and the domain walls are located at $`L_i(i=1,2,\mathrm{},n1)`$. At this stage, one might ask whether it is possible to find a solution with odd domain walls or not. The answer does not seem to be affirmative. The reason is that as in an orbifold geometry $`S^1/Z_2`$ , the solutions obtained in this paper have a characteristic feature that number of positive energy branes equals to number of negative energy branes, so the resulting solutions always include even domain walls. Now let us turn to our main question. ”How is mass hierarchy problem resolved in our many universe model?” We will see that the answer is quite of interest and shows a peculiar feature of many universe model. First of all, let us consider the first model (14). In this model, domain walls are located at $`z=L_{2i}(i=1,2,\mathrm{},\frac{n1}{2})`$. Let us suppose that a visible 3-brane (our universe) is located at a certain $`z=L_{2i}`$, whereas the remaining 3-branes are hidden sectors (the other universes). We wish to measure mass scale in our world placed at $`z=L_{2i}`$ by mass scale in a hidden world placed at $`z=L_{2j}(ij)`$, to which we assume that the Planck mass is assigned. According to formulas given in , we can in general evaluate mass scale in our world by using Eqs.(12), (14): $`m(L_{2i})`$ $`=`$ $`{\displaystyle \frac{u(L_{2i})}{u(L_{2j})}}m(L_{2j})`$ (19) $`=`$ $`m(L_{2j})\mathrm{exp}\left\{kr_c\left[L_{2i}L_{2j}+{\displaystyle \underset{l=1}{\overset{\frac{n1}{2}}{}}}(1)^l\left(|L_{2i}L_{2l}||L_{2j}L_{2l}|\right)\right]\right\}.`$ To compare the result (19) to that of , let us first recall the result of and then see the implications of our result in the two specific cases. In the setup of Ref., there are two 3-branes with opposite sign of potential energy. For a 3-brane with positive potential energy, a natural scale for mass is assumed to be of order the Planck mass. If the Standard Model (our universe) is placed on another 3-brane with negative potential energy, the graviton amplitude is exponentially suppressed and exponential mass hierarchy is generated. To have this fact in mind, let us see what happens in our model. From the action (1) and (16), it turns out that 3-branes located at $`z=L_{4i2}`$ have negative brane energy while 3-branes at $`z=L_{4i}`$ have positive brane energy. Thus, for instance, let us calculate mass scale at $`z=L_2`$ from mass scale at $`z=L_4`$, for which we assume that the Planck mass is assigned. Note that this situation is similar to that of as mentioned above. A simple calculation yields $`m(L_2)=m(L_4)e^{kr_c(L_4L_2)}.`$ (20) This result means that exponential mass hierarchy also occurs in the present model as in . Note that this result is universal in the sense that the essential behavior of the result remains unchanged as long as our 3-brane has negative brane energy and the hidden 3-brane has positive brane energy even if a factor $`L_4L_2`$ in the exponential changes depending on which 3-branes we consider. An interesting feature of our model is the existence of many branes with negative and positive brane energy, so we are now led to ask how the result about mass hierarchy is modified when we take account of two 3-branes with same sign of brane energy. To be specific, let us consider a situation where we attempt to measure mass scale at $`z=L_2`$ from mass scale at $`z=L_6`$, of which both the 3-branes have negative brane energy. In this case the general result (19) reduces to $`m(L_2)=m(L_6)e^{\frac{1}{2}kr_c(L_3+L_5L_7)},`$ (21) which implies that there is also exponential mass hierarchy if $`L_3+L_5>L_7`$. But if this inequality does not hold, then the Planck mass scale allocated in hidden sector is exponentially enhanced to the larger mass scale in our world. It is easy to check that the essential feature of this conclusion is also universal as far as two 3-branes with same sign of brane energy are concerned. Next, let us focus on the second model (17). In this model, domain walls are located at $`z=L_i(i=1,2,\mathrm{},n1)`$. Again, a simple calculation gives us $`m(L_i)`$ $`=`$ $`{\displaystyle \frac{u(L_i)}{u(L_j)}}m(L_j)`$ (22) $`=`$ $`m(L_j)\mathrm{exp}\left\{kr_c{\displaystyle \underset{l=2}{\overset{n1}{}}}(1)^{l+1}\left(|L_iL_l||L_jL_l|\right)\right\}.`$ From the action (1) and (18), we see that 3-branes located at $`z=L_{2i}`$ have negative brane energy while 3-branes at $`z=L_{2i1}`$ have positive brane energy. Thus, as in the first model, for instance, let us calculate mass scale at $`z=L_2`$ from mass scale at $`z=L_1=0`$, for which we take the Planck mass. The result is of the form $`m(L_2)=m(0)e^{kr_cL_2}.`$ (23) Indeed, this result is the same as in . (Recall that for comparison we have to set $`L_2`$ to $`\pi `$.) An interesting thing here is that this holds true even if there are many hidden sectors and the topology is a circle $`S^1`$ instead of an orbifold $`S^1/Z_2`$. Therefore, again we have exponential mass hierarchy even in the second model. Moreover, it is straightforward to show that even if we take account of two 3-branes with same sign of brane energy we have also exponential mass hierarchy if a suitable inequality holds like the first model. Finally, let us consider mass scale of branes at $`z=L_{2i1}`$. Since we have an equation $`u(L_{2i1})=1`$, the branes at these locations have special feature. Namely, if we attempt to evaluate mass scale at $`z=L_{2i1}`$ from mass scale at $`z=L_{2j1}`$ with $`ij`$, there is no mass hierarchy, in other words, $`m(L_{2i1})=m(L_{2j1})`$. However, this is an artifact of the model. In fact, we can modify the value of $`u(L_{2i1})`$ from one to a non-zero constant without violating the validity of the Einstein equation, so in such a situation we have exponential mass hierarcy as before. To summarize, we have investigated a possibility of constructing a new model with an exponential mass hierarchy whose existence is inspired by the perspective of D-brane theory and many universe cosmology. In our model, the fifth dimension has a topology $`S^1`$ rather than a singular orbifold $`S^1/Z_2`$, which is one of advantages in our model. It was shown that even in the present model, we have the exponential mass hierarchy under an appropriate condition. However, there are many universes in the present model, perhaps our universe and the other many hidden universes, so there is $`a\mathrm{𝑝𝑟𝑖𝑜𝑟𝑖}`$ no way of determining which universe our universe is among many universes. Hence, the correct interpretation of our results is the following: mass scale in our universe depends on a relative distance relation along the fifth dimension between our universe and hidden universes. Maybe once God created many universes and gave the Planck scale to one (or some) universe(s) in the beginning of the big bang, mass scales in various universes are fixed by a relative relation among universes. According to this new scenario, each universe should have $`\mathrm{𝑛𝑜𝑡}`$ more than one $`\mathrm{𝑏𝑢𝑡}`$ exactly one low (or high) energy scale, and our world happens to have taken the electro-weak scale as such a low energy scale. Although we have so far presented only two types of solutions which satisfy the Einstein equation and have simple and manageable forms, we have also examined the other solutions in some detail . The conclusion is almost the same as in the present cases, so we think that the two solutions account for the essential features in our theory. One disadvantage of our model as well as the Randall-Sundrum original model is the existence of 3-branes with negative potential energy. Recently, an alternative setup has been put forward where only positive energy objects are taken into account . We wish to investigate a possibility of generalizing our model to such a direction in future. Acknowledgement We are indebted to M. Tonin for stimulating discussions and continuous encouragement. We wish to thank Dipartimento Di Fisica, ”Galileo Galilei”, Universita Degli Studi Di Padova, for a kind hospitality, where most of parts of this work have been done.
no-problem/9908/nucl-th9908035.html
ar5iv
text
# Low–energy 𝜂⁢𝑑 resonance ## Abstract Elastic $`\eta d`$ scattering is considered within the Alt–Grassberger–Sandhas (AGS) formalism for various $`\eta N`$ input data. A three–body resonant state is found close to the $`\eta d`$ threshold. This resonance is sustained for different choices of the two–body $`\eta N`$ scattering length $`a_{\eta N}`$. The position of the resonance moves towards the $`\eta d`$ threshold when $`\mathrm{Re}a_{\eta N}`$ is increased, and turns into a quasi-bound state at $`\mathrm{Re}a_{\eta N}=0.733`$ fm. PACS numbers: 25.80.-e, 21.45.+v, 25.10.+s One of the main questions of $`\eta `$–meson physics concerns the existence of $`\eta `$–nuclei, a possibility predicted in a pioneering work by Haider and Liu . In this context it is worth mentioning that according to the mean free path of $`\eta `$ mesons in a nuclear medium is about 2 fm, i.e., less than the size of a typical nucleus. A necessary condition for the existence of $`\eta `$–nuclei, hence, appears to be satisfied. A further indication is the observation of pionic nuclei (deeply bound pion states inside nuclei) , which were predicted theoretically in . Recently, in a number of theoretical investigations based on quite different approaches, such as the mean field approximation , the optical model , and few–body calculations , the idea of $`\eta `$–nuclei was given a rather firm ground. Experimentally, after the discouraging attempt of Ref. , certain evidence for the existence of $`\eta `$–nuclei was noticed by two groups . An additional indication is given by the enhancement of near–threshold $`\eta `$ production in the reaction $`npd\eta `$ as reported by the Uppsala group . This observation is most likely to be understood as the effect of a near–threshold bound or resonant $`\eta d`$ state. Moreover, it was suggested that even quite exotic systems containing an $`\eta `$ meson and a hyperon ($`\eta `$–hypernuclei) may exist. Being of interest by itself, the existence of $`\eta `$–nuclei would also shed new light on various fundamental problems of particle physics. For example, the study of $`\eta `$–nuclei would give a clue for understanding the possible restoration of chiral symmetry in a nuclear medium, or its partial restoration which may occur at normal densities $`0.17\mathrm{fm}^3`$ . A further aspect, which makes such systems interesting for other fields of nuclear physics, is the modification of the two–body $`\eta N`$ force by the nuclear medium. Being enhanced due to the resonant character of the $`\eta N`$ interaction, general properties of such modifications should become particularly evident in this case. Especially, the structure of the $`S_{11}`$ resonance embedded in a nuclear medium may be studied. This would exhibit certain details of the effective Lagrangian models of such a resonance , and shed some light, for example, on chiral models which suggest a reduction of its partial width for decaying into the $`\eta N`$ channel , or assume $`S_{11}`$ to be a quasi-bound state of the $`K`$ meson and a $`\mathrm{\Sigma }`$ hyperon . The interesting suggestion of considering $`S_{11}`$ as a joint manifestation of threshold (casp) and resonance phenomena could also be checked. In a sense, the problem of choosing the most adequate model for describing the $`S_{11}`$ resonance is similar to the usual difficulty of choosing the “correct” two–body potential. A reliable judgment on such a choice can be given when this potential (or the resonance) is used in a few–body system where its off-shell properties are exposed. This requires a treatment based on rigorous few–body theory. The Alt–Grassberger–Sandhas (AGS) equations , which are employed in this work to calculate the $`\eta `$–deuteron elastic scattering amplitude, belong to this category. The Faddeev–type coupling of these equations guarantees uniqueness of their solutions. Moreover, as equations for the elastic and rearrangement operators $`U_{\alpha \alpha }`$ and $`U_{\beta \alpha }`$, respectively, they are well–defined in momentum space, providing thus the desired scattering amplitudes in a most direct and technically reliable manner. The advantage of working with coupled equations is not only suggested by questions of uniqueness, but also by the relevance of rescattering effects. Indeed, in recent calculations of $`\eta `$–photoproduction from the deuteron it was found that rescattering terms give a significant contribution to the corresponding amplitude . Let the $`\eta `$ meson be denoted as particle 1 and the two nucleons as particles 2 and 3. With the momentum $`𝐪_1`$ of the meson relative to the center of mass of the nucleons in the deuteron state $`|\psi _d`$, the free $`\eta d`$ channel state is given by $$|\psi _d;𝐪_1=|\psi _d|𝐪_1,$$ (1) its normalization being chosen as $$\psi _d;𝐪_1^{}|\psi _d;𝐪_1=\delta (𝐪_1^{}𝐪_1).$$ (2) In terms of the AGS transition operator $`U_{11}`$ the $`\eta d`$ elastic scattering amplitude is represented as $$f(𝐪_1^{},𝐪_1)=(2\pi )^2M_1\psi _d;𝐪_1^{}|U_{11}(z)|\psi _d;𝐪_1,$$ (3) with the on–energy–shell conditions $`|𝐪_1^{}|=|𝐪_1|`$ and $`z=E_d+q_1^2/2M_1+i0`$. Here $`M_1`$ is the reduced mass of particle 1 and the (23) subsystem, $`1/M_1=1/m_1+1/(m_2+m_3)`$. The elastic transition operator $`U_{11}`$ satisfies the set of AGS equations $$U_{\beta \alpha }(z)=(1\delta _{\beta \alpha })G_0^1(z)+\underset{\gamma =1}{\overset{3}{}}(1\delta _{\beta \gamma })T_\gamma (z)G_0(z)U_{\gamma \alpha }(z),$$ (4) with $`G_0(z)`$ being the free resolvent (Green’s operator) of the three particles involved. This set of equations couples all $`3\times 3`$ elastic and rearrangement operators $`U_{\alpha \alpha }`$ and $`U_{\beta \alpha }`$. Here each of the subscripts runs through the values 1,2 and 3, indicating the two–fragment partitions (1,23), (2,31) and (3,12) respectively. Therefore, $`U_{11}`$ describes the elastic transition 1(23)$``$1(23), while $`U_{21}`$ represents the rearrangement process 1(23)$``$2(13). In accordance with Eq.(1), we denote by $`𝐪_\alpha `$ the momentum of particle $`\alpha `$ relative to the center of mass of the ($`\beta \gamma `$) subsystem, and by $`M_\alpha `$ the corresponding reduced mass. The internal momentum of this subsystem is denoted as $`𝐩_\alpha `$. Conventionally, this complementary notation is also used to label the two–body $`T`$–operator $`t_\alpha (z)`$ of the ($`\beta \gamma `$) pair, for instance $`t_1(z)=t_{NN}(z)`$. It should be noticed, however, that it is not this genuine two–body operator which enters the AGS equations, but the operator $$T_\alpha (z)=t_\alpha (z\widehat{𝐪}_\alpha ^2/2M_\alpha ),$$ (5) which is to be understood as the two–body operator embedded in the three–body space, with the relative kinetic energy operator $`\widehat{𝐪}_\alpha ^2/2M_\alpha `$ of particle $`\alpha `$ being subtracted from the total energy variable $`z`$. Considered in momentum space, Eq.(5) thus reads $$𝐩_\alpha ^{},𝐪_\alpha ^{}|T_\alpha (z)|𝐩_\alpha ,𝐪_\alpha =\delta (𝐪_\alpha ^{}𝐪_\alpha )𝐩_\alpha ^{}|t_\alpha (zq_\alpha ^2/2M_\alpha )|𝐩_\alpha .$$ (6) Since we are interested in $`\eta d`$ collision the subscript $`\alpha `$ is fixed to 1. Instead of the 9 equations of system (4) we therefore have to consider only the three equations $`U_{11}(z)`$ $`=`$ $`T_2(z)G_0(z)U_{21}(z)+T_3(z)G_0(z)U_{31}(z),`$ (7) $`U_{21}(z)`$ $`=`$ $`G_0^1(z)+T_1(z)G_0(z)U_{11}(z)+T_3(z)G_0(z)U_{31}(z),`$ (8) $`U_{31}(z)`$ $`=`$ $`G_0^1(z)+T_1(z)G_0(z)U_{11}(z)+T_2(z)G_0(z)U_{21}(z),`$ (9) involving the operator $`U_{11}`$ which determines the elastic amplitude (3). In momentum space the system (8) consists, after partial wave decomposition, of three coupled two–dimensional integral equations. It is customary to reduce the dimension of these equations by approximating or representing the two–body $`T`$–operators in separable form, $$t_\alpha (z)=|\chi _\alpha \tau _\alpha (z)\chi _\alpha |,\alpha =1,2,3,$$ (10) or, according to (5), by $$T_\alpha (z)=|\chi _\alpha \tau _\alpha (z\widehat{𝐪}_\alpha ^2/2M_\alpha )\chi _\alpha |,\alpha =1,2,3.$$ (11) Inserting this representation in the AGS equations (4) and sandwiching them between $`\chi _\beta |`$ and $`|\chi _\alpha `$ they take the form $$X_{\beta \alpha }(z)=(1\delta _{\beta \alpha })\chi _\beta |G_0(z)|\chi _\alpha +\underset{\gamma =1}{\overset{3}{}}(1\delta _{\beta \gamma })\chi _\beta |G_0(z)|\chi _\gamma \tau _\gamma (z\frac{\widehat{𝐪}_\gamma ^2}{2M_\gamma })X_{\gamma \alpha }(z).$$ (12) The three–body operators $`U_{\beta \alpha }`$, hence, are replaced by the effective two–body operators $$X_{\beta \alpha }(z)=\chi _\beta |G_0(z)U_{\beta \alpha }(z)G_0(z)|\chi _\alpha $$ (13) which act exclusively on the relative momentum states $`|𝐪_\alpha `$. After partial wave decomposition we then end up with one-dimensional integral equations. The form-factor $`|\chi _1`$ in the $`NN`$ operator $`t_1`$ is related to the deuteron wave function $`|\psi _d`$ according to $$|\psi _d;𝐪_1=G_0\left(q_1^2/2M_1+E_d\right)|\chi _1|𝐪_1,$$ (14) where $`E_d`$ is the deuteron energy. The on-energy-shell matrix element in (3), thus, can be written in the form $$f(𝐪_1^{},𝐪_1)=(2\pi )^2M_1𝐪_1^{}|X_{11}(z)|𝐪_1.$$ (15) In other words, the set of equations (12) provides the elastic amplitude we are interested in. In the present case there are further simplifications caused by the identity of the nucleons. Indeed, the momentum representation of $`X_{31}`$, $`\tau _3`$, $`|\chi _3`$ and of $`X_{21}`$, $`\tau _2`$, $`|\chi _2`$ respectively are of the same functional form. This reduces the three equations involving $`X_{11}`$ to the following pair $`X_{11}(z)`$ $`=`$ $`2\chi _1|G_0(z)|\chi _2\tau _2\left(z\widehat{𝐪}_2^2/2M_2\right)X_{21}(z),`$ (16) $`X_{21}(z)`$ $`=`$ $`\chi _2|G_0(z)|\chi _1+\chi _2|G_0(z)|\chi _1\tau _1\left(z\widehat{𝐪}_1^2/2M_1\right)X_{11}(z)`$ (17) $`+\chi _2|G_0(z)|\chi _2\tau _2\left(z\widehat{𝐪}_2^2/2M_2\right)X_{21}(z).`$ (18) The $`S`$–wave nucleon–nucleon separable potential is adopted from Ref. with its parameters slightly modified to be consistent with more recent $`NN`$ data (see Ref.). The $`\eta `$–nucleon $`T`$–matrix is taken in the form $$t_{\eta N}(p^{},p;z)=(p_{}^{}{}_{}{}^{2}+\alpha ^2)^1\frac{\lambda }{(zE_0+i\mathrm{\Gamma }/2)}(p^2+\alpha ^2)^1$$ (19) consisting of two vertex functions and the $`S_{11}`$-propagator in between . It corresponds to the process $`\eta NS_{11}\eta N`$ which at low energies is dominant. The range parameter $`\alpha =3.316`$ fm<sup>-1</sup> was determined in Ref. , while $`E_0`$ and $`\mathrm{\Gamma }`$ are the parameters of the $`S_{11}`$ resonance , $$E_0=1535\mathrm{MeV}(m_N+m_\eta ),\mathrm{\Gamma }=150\mathrm{MeV}.$$ The strength parameter $`\lambda `$ is chosen to reproduce the $`\eta `$–nucleon scattering length $`a_{\eta N}`$, $$\lambda =\frac{\alpha ^4(E_0i\mathrm{\Gamma }/2)}{(2\pi )^2\mu _{\eta N}}a_{\eta N}.$$ (20) It is customary to use complex $`a_{\eta N}`$ the imaginary part of which accounts for the flux losses into the $`\pi N`$ channel. The value of $`a_{\eta N}`$ is not accurately known. Different analyses provided for $`a_{\eta N}`$ values in the range $$0.27\mathrm{fm}\mathrm{Re}a_{\eta N}0.98\mathrm{fm},0.19\mathrm{fm}\mathrm{Im}a_{\eta N}0.37\mathrm{fm}.$$ (21) Recently, however, most of the authors agreed that $`\mathrm{Im}a_{\eta N}`$ is around 0.3 fm. But for $`\mathrm{Re}a_{\eta N}`$ the estimates are still very different (compare, for example, Refs. and ). We, therefore, fixed $`\mathrm{Im}a_{\eta N}`$ to 0.3 fm and did calculations for several values of $`\mathrm{Re}a_{\eta N}`$ within the above interval. We solved Eqs. (17) for $`\eta d`$ collision energies varying from zero ($`\eta d`$-threshold, $`z=E_d`$) up to 22 MeV. As is well known (see, for example, ), the kernels of Eqs.(17), when expressed in momentum representation, have logarithmic singularities for $`z>0`$. These singularities stem from the Green’s functions which in their denominators involve two terms $`q_\alpha ^2/2M_\alpha `$ and $`p_\alpha ^2/2\mu _\alpha `$ corresponding to the pair of Jacobi momenta. The singularities appear after angular integration, and their position depends on the values of $`p_\alpha `$ and $`q_\alpha `$ (the so-called moving singularities). In the numerical procedure, we handle it with the method suggested in Ref. . The main idea of this method consists in expanding the unknown solutions (in the area covering the singular points) in certain polynomials and subsequent analytic integration of the singular part of the kernels. The results of our calculations are presented in Figs. 15 and in Table I. In Figs. 1 and 2 the energy dependence of the $`\eta d`$ phase-shifts for five different choices of $`\mathrm{Re}a_{\eta N}`$ is shown, namely, for 0.55 fm, 0.65 fm, 0.725 fm, 0.75 fm, and 0.85 fm. The larger this value, the stronger is the $`\eta `$N attraction. The change in the character of these curves, hence, reflects the growth of the attractive force between the $`\eta `$ meson and the nucleon. The lower three curves for $`\mathrm{Re}\delta _{\eta d}`$ corresponding to the smaller values of $`\mathrm{Re}a_{\eta N}`$ start from zero, the two curves corresponding to the strong attraction start from $`\pi `$. According to Levinson’s theorem, the phase shift at threshold energy is equal to the number of bound states $`n`$ times $`\pi `$. We found that the transition from the lower family of the curves to the upper one happens at $`\mathrm{Re}a_{\eta N}`$=0.733 fm. Therefore, the $`\eta `$N force, which generates $`\mathrm{Re}a_{\eta N}>0.733`$ fm, is sufficiently attractive to bind $`\eta `$ inside the deuteron. This is the first conclusion of our calculations. The second conclusion concerns the peaks in the energy dependence of the total elastic cross-section (see Fig. 4), indicating that a resonance appears in the $`\eta d`$–system. Of course, not every maximum of the cross-section is a resonance, but the Argand plots, shown in Fig. 5, prove that the maxima we found are resonances. Their positions and widths for various choices of $`\mathrm{Re}a_{\eta N}`$ are given in Table I. It should be noted that, while the resonance energy is determined in our calculations exactly (as the maximum of the function $`\mathrm{sin}^2\mathrm{Re}\delta _{\eta d}`$), the corresponding width is obtained by fitting the cross-section with a Breit–Wigner curve. Therefore, the values of $`\mathrm{\Gamma }_{\eta d}`$ given in Table I should be considered only as rough estimates. The presence of a resonance before a quasi-bound state appears is not surprising. With increasing attraction the poles of the $`S`$-matrix should move in the complex plane from the resonance area to the quasi-bound state area. This is exactly what our calculations indicate. In Ref. we showed that such transition of the pole happens when $`\mathrm{Re}a_{\eta N}`$ changes from 0.25 fm to 1 fm. Here we found more exactly that it happens at $`\mathrm{Re}a_{\eta N}`$=0.733 fm. The resonant behaviour of $`\eta d`$ elastic scattering should be seen in various processes involving $`\eta d`$–system in their final states, such as $`\gamma dd\eta `$ and $`npd\eta `$. Recent measurements of the $`\eta `$–production in the $`np`$ collisions reveal a bump of the cross–section for the reaction $`npd\eta `$ at a c.m. energy below 5 MeV. If we suppose that the energy dependence of this cross–section is mainly determined by the final state interaction, then this bump can be explained by the existence of an $`\eta d`$ resonance. Moreover, since the bump was observed below 5 MeV, the resonance positions given in Table I imply the rough lower bound $`\mathrm{Re}a_{\eta N}0.75\mathrm{fm}`$. For a reliable estimate, however, one has to perform an explicit calculation of the corresponding cross-section. A recent analysis of this reaction with a non-resonant final state interaction is consistent with the data of Ref. only if $`\mathrm{Re}a_{\eta N}=0.30\mathrm{fm}`$. Finally, it is interesting to note that there is tentative evidence for a similar bump in the low–energy region in the reaction $`\gamma dX\eta `$ . ###### Acknowledgements. Financial support from Deutsche Forschungsgemeinschaft, NATO (grant #CRGLG970110), INTAS (grant #96-0457), International Center for Fundamental Physics (Moscow), Russian Foundation for Basic Research, National Research Foundation of South Africa and the Joint Institute for Nuclear Research (Dubna) is greatly appreciated.
no-problem/9908/chao-dyn9908003.html
ar5iv
text
# Synchronisation of time–delay systems ## Abstract We present the linear-stability analysis of synchronised states in coupled time–delay systems. There exists a synchronisation threshold, for which we derive upper bounds, which does not depend on the delay time. We prove that at least for scalar time–delay systems synchronisation is achieved by transmitting a single scalar signal, even if the synchronised solution is given by a high–dimensional chaotic state with a large number of positive Lyapunov–exponents. The analytical results are compared with numerical simulations of two coupled Mackey–Glass equations. The problem of synchronisation of dynamical systems is one of the classical fields in engineering science . Recently, renewed interest in this field was stimulated in connection with the synchronisation of chaotic motion. Especially, the potential applicability for communication has attracted much research in recent years . Yet, there are a lot of results available concerning the synchronisation of low–dimensional chaotic systems, theoretical as well as experimental . Contrary, the synchronisation of high–dimensional chaotic systems with possibly a large number of positive Lyapunov–exponents remains open. From the point of view of numerical simulations the synchronisation of specific high–dimensional chaotic systems has been achieved , while, to our best knowledge, rigorous results, e. g. concerning sufficient synchronisation conditions, are still lacking. While it has been proved recently by Stojanovski et al. that the synchronisation of high-dimensional chaotic states can be in principle achieved with a single transmitted variable, the problem of finding the appropriate coupling of the two systems remains open. For that reason we address in this paper the question of the synchronisation of coupled identical time–delay systems. We focus on time–delay systems, since on the one hand it is well established that these systems are prominent examples of high–dimensional chaotic motion with a large number of positive Lyapunov–exponents , and one the other hand synchronisation of Mackey–Glass type electronic oscillators has been reported from the experimental point of view . Let us consider a fairly general theoretical model and investigate the stability problem of a synchronised state. For that purpose consider two identical arbitrary scalar time–delay systems with a symmetric coupling $`\dot{x}`$ $`=`$ $`F(x,x_\tau )K(xy),`$ (1) $`\dot{y}`$ $`=`$ $`F(y,y_\tau )K(yx),`$ (2) where we adopt the notation $`x_\tau :=x(t\tau )`$ to indicate the time–delayed variables. We specialise from the beginning to the frequently analysed case that the coupling is bi–directional and acts additive to the single dynamical system. However, we stress that the subsequent considerations apply with minor modifications to much more general situations, e. g. to vector–type variables, to systems with much more general delay terms, or to a non–additive coupling, as long as the coupling vanishes in the synchronised state $`x(t)y(t)`$. But we think, that the choice made in eq.(1) makes our arguments more transparent. Let $`z`$ denote the synchronised solution, i.e. $`\dot{z}=F(z,z_\tau )`$. Considering deviations from that state according to $`x=z+\delta x`$, $`y=z+\delta y`$ and performing a linear stability analysis, we obtain for the deviation $`\mathrm{\Delta }:=\delta y\delta x`$ from the synchronised state the linear differential–difference equation $$\dot{\mathrm{\Delta }}=\alpha (t)\mathrm{\Delta }+\beta (t)\mathrm{\Delta }_\tau .$$ (3) Here, the time–dependent coefficients are given in terms of the synchronised solution as $`\alpha (t)=_1F(z,z_\tau )2K`$ and $`\beta (t)=_2F(z,z_\tau )`$, where the symbol $`_{1/2}`$ denotes the derivative with respect to the first/second argument. A superficial inspection of eq.(3) might suggest that the synchronised solution is stable if $`\alpha (t)`$ is ”sufficiently negative”. In fact, we will make this statement rigorous in what follows. Suppose the coefficients are bounded in the sense that $`\alpha (t)a<0`$ and $`|\beta (t)|b`$ holds for some fixed values $`a`$ and $`b`$. Since the equation is linear, it is sufficient to analyse the solution with the special initial condition $`\mathrm{\Delta }(0)=1`$, $`\mathrm{\Delta }(t)0,t<0`$. The general case follows by a simple integration. There are different ways to estimate the stability of the trivial solution, $`\mathrm{\Delta }(t)0`$, of eq.(3). Here we use the fact that for scalar quantities a simple closed analytical formula for the solution can be written down. One just integrates the linear equation (3) in the time intervals $`[N\tau ,(N+1)\tau ]`$ and considers the delay term as an inhomogeneous part. By this continuation method (cf.\[8, p.45\]) the full solution is obtained as $`\mathrm{\Delta }(t)`$ $`=`$ $`e^{_0^t\alpha (t^{})𝑑t^{}}+{\displaystyle _\tau ^t}𝑑t_1\beta (t_1)e^{_{I_1}\alpha (\theta )𝑑\theta }`$ (4) $`+`$ $`{\displaystyle _{2\tau }^t}𝑑t_1{\displaystyle _{2\tau }^{t_1}}𝑑t_2\beta (t_1)\beta (t_2\tau )e^{_{I_2}\alpha (\theta )𝑑\theta }+\mathrm{}`$ (5) $`+`$ $`{\displaystyle _{N\tau }^t}𝑑t_1{\displaystyle _{N\tau }^{t_1}}𝑑t_2\mathrm{}{\displaystyle _{N\tau }^{t_{N1}}}𝑑t_N`$ (8) $`\beta (t_1)\beta (t_2\tau )\mathrm{}\beta (t_N(N1)\tau )e^{_{I_N}\alpha (\theta )𝑑\theta }`$ $`\text{for }N\tau t(N+1)\tau .`$ Here, the domains of integration for the exponents are given by $`I_k:=[0,t]/([t_1,t_1\tau ][t_2\tau ,t_22\tau ]\mathrm{}[t_k(k1)\tau ,t_kk\tau ])`$. An upper bound for $`|\mathrm{\Delta }(t)|`$ is obtained, if the maximal values $`\alpha (t)=a`$ and $`\beta (t)=b`$ are inserted into eq.(4). But then, the expression reduces to a solution $`\mathrm{\Gamma }`$ of the differential–difference equation with constant coefficients $$\dot{\mathrm{\Gamma }}=a\mathrm{\Gamma }+b\mathrm{\Gamma }_\tau .$$ (9) Hence, a solution of eq.(9) yields an upper bond for $`|\mathrm{\Delta }(t)|`$. But the last equation is easily solved by a Laplace transformation (cf.) or loosely speaking by an exponential ansatz $`\mathrm{\Gamma }(t)=e^{st}`$. Since the corresponding eigenvalues obey $`s=a+b\mathrm{exp}(s\tau )`$, negative real parts, i. e. stability, occur if and only if $`a>b`$. This inequality yields an upper bond $`K_+`$ for the critical coupling strength beyond which synchronisation is achieved. If we take the definitions of $`\alpha (t)`$ and $`\beta (t)`$ into account it reads explicitely $$K_+=1/2[\underset{t}{\mathrm{max}}_1F(z,z_\tau )+\underset{t}{\mathrm{max}}|_2F(z,z_\tau )|].$$ (10) We note as a by–product that eq.(9) may be viewed as a kind of Grownwall–like lemma for the time–dependent equation (3). In what follows, we compare our analytical result to numerical simulations. We specialise to the Mackey–Glass system, i. e. $$F(x,x_\tau )=x+\frac{ax_\tau }{1+x_\tau ^{10}}.$$ (11) In order to investigate the properties of the synchronisation mechanism by numerical methods, we chose the distance between trajectories as a suitable measure. For that reason the quantity $$D_T(t)=_t^{t+T}x(t^{})y(t^{})𝑑t^{},$$ (12) which of course depends on the range of averaging $`T`$ and the point of reference $`t`$ was analysed. We used a Runge–Kutta algorithm of fourth order with step size $`0.1`$. The simulation, which have been performed for the parameter value $`a=3`$, started with $`K=0.35`$. A constant initial condition for $`x`$ and $`y`$, which differs by an amount of $`10^3`$ has been chosen. The system was allowed to relax for a time $`t=80\tau `$. After that, the distance $`D`$ was integrated on a trajectory of the length $`T=80\tau `$. For the next value of coupling strength $`K`$ we again distorted the last state of the trajectory by adding an amount of $`10^3`$ to the $`y`$–coordinates and used it as initial condition. We performed the computation for increasing as well as decreasing coupling constant. Fig.1 summarises our findings. For $`\tau =10`$ we observe distinct jumps, indicating that the system switches between coexisting periodic states with a pronounced hysteresis. For $`\tau =100`$, the overall behaviour of the system appears to be quite similar as for $`\tau =10`$, except that no switching and no hysteresis is observed. Within the resolution of the graphics the same behaviour has already been observed for a smaller value $`\tau =50`$. From the numerical simulations the synchronisation threshold is estimated as $`K_c(\tau =10)0.24`$, $`K_c(\tau =100)0.28`$. If we evaluate our analytical estimate eqs.(10) and (11) using upper bounds for the derivatives we obtain values which differ by an order of magnitude but are independent of the delay time $`\tau `$, $`K_+=(81a/401)/2=2.5375`$. Since we have applied a rather graceful rigorous estimate such a discrepancy is far from being astonishing. In order to understand the dynamics in the vicinity of the synchronisation threshold, time traces of the difference $`x(t)y(t)`$ have been computed (cf. fig.2). Slightly below the synchronisation threshold $`K_c`$ we observe an intermittent behaviour very similar to on–off intermittency . Additionally, we investigated the distribution of laminar phases and turbulent phases under variation of the coupling strength $`K`$ and the delay time $`\tau `$, which we present in fig. 3. To this end, the distance $`D_\tau (t)`$ of the two systems in the phase space has been computed on an intervall of length $`\tau 10^6`$. Then, the length of the laminar phases ($`D_\tau 0.10`$) and turbulent phases ($`D_\tau >0.10`$) were recorded. We observe a power-law scaling of the distribution $`P_l`$ of the laminar phases over a wide range, $`P_lt^{\alpha _l(\tau )}`$, where the exponent $`\alpha _l`$ depends slightly on the delay time. In the low-dimensional chaotic case, for $`\tau =10.0`$, we observe $`\alpha (10)=1.50`$ in agreement with the value predicted by the scenario of on-off intermittency. In the high-dimensional chaotic cases, for increasing delay time, we observe a decreasing exponent: $`\alpha (30.0)=1.41,\alpha (50.0)=1.38,\alpha (100.0)=1.27`$, indication that there might be deviations from on–off intermittency. The distribution of the turbulent phases follow a $`P_tt^{\alpha _t(\tau )}`$-scaling for high enough $`\tau `$. Although these findings are at the first sight not surprising, a detailed analysis is required, with special focus on the high dimensionality of the dynamics. Details will be reported elsewhere. We conclude with a remark, how our results depend on noise or other imperfections which are present in realistic systems. In fact, in order to apply a concept like synchronisation such perturbations have to be small and we may assume a general linear dependence. Formally such contributions are introduced into eq.(1) by adding the two terms $`G(x,x_\tau )\xi `$ and $`G(y,y_\tau )\eta `$, where $`\xi `$ and $`\eta `$ denote for example realisations of a noise. Considering the perturbations of the same order of magnitude like the deviation from the unperturbed synchronised state and proceeding as above we finally end up with $$\dot{\mathrm{\Delta }}=\alpha (t)\mathrm{\Delta }+\beta (t)\mathrm{\Delta }_\tau +G(z,z_\tau )(\eta \xi ),$$ (13) which differs from eq.(3) just by an inhomogeneous contribution. The theory of linear difference–differential equations tells us that eq.(13) inherits its stability properties from the corresponding homogeneous system (3) except that the perturbations cause fluctuations around the unperturbed synchronised state. Whenever the perturbations are so large, that contributions beyond the linear order have to be taken into account, one has to resort to different methods. One of these cases, which are also relevant from the experimental point of view, is given by the synchronisation of nearly-identical time-delay systems. Since, in this case, no strict synchronised solution $`xy`$ exists, one has to rely on more general concepts, such as the generalised synchronisation . In summary, we emphasise that an analytical upper bond for the solution of eq.(3) is obtained if one replaces the time dependency of the coefficients by their extreme values. One might get better estimates in special cases. In particular one might argue, that eq.(9) already determines the stability, if the time averages of the coefficients are inserted. This statement is in fact true if either the coefficients are periodic functions of time with the delay $`\tau `$ being an integer multiple of the period or, if the coefficients are almost constant (cf.\[8, p.277\]). Whether the general case can be treated by this refined estimate remains open. Nevertheless we have shown, that for sufficiently large coupling constant $`K`$ the synchronised solution of eq.(1) becomes stable, whenever $`|\delta _{1/2}F(z,z_\tau )|`$ are uniformly bounded. In particular the critical coupling strength remains bounded even in the limit of large delay times. i. e. it does not increase with the dimension of the attractor. In fact, our numerical simulations indicate only a weak dependence of the actual critical coupling strength on the delay time. Last but not least our approach clearly demonstrates that the success of the synchronisation is independent of the number of positive Lyapunov–exponents, even if our coupling uses one scalar variable only, illustrating the results of Stojanovski et al. . We acknowledge discussions with U. Parlitz, K. Pyragas, and Th. Meyer.
no-problem/9908/hep-ph9908345.html
ar5iv
text
# 1. Introduction ## 1. Introduction A lot of data have been accumulated on the top-quark since its discovery. However it is still an open question whether its interactions obey the standard scheme like all the other fermions or there exists some new-physics contribution to its couplings. The top quark decays immediately after being produced because of its huge mass. Therefore the decay process is not influenced by any hadronization effects and consequently its decay products are expected to carry valuable information on the top properties. Next linear colliders (NLC) of $`e\overline{e}`$ will give us fruitful data on the top through $`e\overline{e}t\overline{t}`$. In particular the energy spectra of the lepton(s) produced in its semileptonic decay(s) turn out to be a useful analyzer of the top-quark couplings . Indeed many authors have worked on this subject (see the reference list of Ref.), and we also have tackled them over the past several years. Here I would like to show some of the results of our latest model-independent analyses via arbitrary longitudinal beam polarizations, where we have assumed the most general anomalous couplings both in the production and decay vertices in contrast to most of the existing works. ## 2. Framework We can represent the most general $`t\overline{t}`$ couplings to the photon and $`Z`$ boson as $$\mathrm{\Gamma }_{vt\overline{t}}^\mu =\frac{g}{2}\overline{u}(p_t)\left[\gamma ^\mu \{A_v+\delta A_v(B_v+\delta B_v)\gamma _5\}+\frac{(p_tp_{\overline{t}})^\mu }{2m_t}(\delta C_v\delta D_v\gamma _5)\right]v(p_{\overline{t}})$$ (1) in the $`m_e=0`$ limit, where $`g`$ denotes the $`SU(2)`$ gauge coupling constant, $`v=\gamma ,Z`$, and $$A_\gamma =\frac{4}{3}\mathrm{sin}\theta _W,B_\gamma =0,A_Z=\frac{v_t}{2\mathrm{cos}\theta _W},B_Z=\frac{1}{2\mathrm{cos}\theta _W}$$ with $`v_t1(8/3)\mathrm{sin}^2\theta _W`$. Among the above form factors, $`\delta A_{\gamma ,Z}`$, $`\delta B_{\gamma ,Z}`$, $`\delta C_{\gamma ,Z}`$ and $`\delta D_{\gamma ,Z}`$ are parameterizing $`CP`$-conserving and $`CP`$-violating non-standard interactions, respectively. On the other hand, we adopted the following parameterization of the $`Wtb`$ vertex suitable for the $`tW^+b`$ and $`\overline{t}W^{}\overline{b}`$ decays: $`\mathrm{\Gamma }_{Wtb}^\mu ={\displaystyle \frac{g}{\sqrt{2}}}\overline{u}(p_b)\left[\gamma ^\mu (f_1^LP_L+f_1^RP_R){\displaystyle \frac{i\sigma ^{\mu \nu }k_\nu }{M_W}}(f_2^LP_L+f_2^RP_R)\right]u(p_t),`$ (2) $`\overline{\mathrm{\Gamma }}_{Wtb}^\mu ={\displaystyle \frac{g}{\sqrt{2}}}\overline{v}(p_{\overline{t}})\left[\gamma ^\mu (\overline{f}_1^LP_L+\overline{f}_1^RP_R){\displaystyle \frac{i\sigma ^{\mu \nu }k_\nu }{M_W}}(\overline{f}_2^LP_L+\overline{f}_2^RP_R)\right]v(p_{\overline{b}}),`$ (3) where $`k`$ is the momentum of $`W`$ and $`P_{L/R}=(1\gamma _5)/2`$. It is worth to mention that the form factors for top and anti-top satisfy the following relations : $$f_1^{L,R}=\pm \overline{f}_1^{L,R},f_2^{L,R}=\pm \overline{f}_2^{R,L},$$ (4) where upper (lower) signs are those for $`CP`$-conserving (-violating) contributions. For the initial beam-polarization we used the following convention: $`P_e^{}=+[N(e^{},+1)N(e^{},1)]/[N(e^{},+1)+N(e^{},1)],`$ (5) $`P_{e^+}=[N(e^+,+1)N(e^+,1)]/[N(e^+,+1)+N(e^+,1)],`$ (6) where $`N(e^{(+)},h)`$ is the number of $`e^{}(e^+)`$ with helicity $`h`$ in each beam. ## 3. Lepton-energy spectra After some calculations, we arrived at the normalized single distribution, which we express as $$\frac{1}{\sigma }\frac{d\sigma }{dx}=\underset{i=1}{\overset{3}{}}c_i^\pm f_i(x).$$ (7) Here $`\pm `$ corresponds to $`\mathrm{}^\pm `$ and the variable $`x`$ is defined from the top velocity $`\beta `$ and the lepton energy $`E_{\mathrm{}}`$, both in the $`e\overline{e}`$ c.m. frame, as $$x\frac{2E_{\mathrm{}}}{m_t}\left(\frac{1\beta }{1+\beta }\right)^{1/2}.$$ The coefficients $`c_i^\pm `$ on the right-hand side are given by $$c_1^\pm =1,$$ $$c_2^\pm =a_1\delta D_V^{()}a_2[\delta D_A^{()}\mathrm{Re}(G_1^{()})]+a_3\mathrm{Re}(\delta D_{VA}^{()})\xi ^{()},$$ $$c_3^+=\mathrm{Re}(f_2^R),c_3^{}=\mathrm{Re}(\overline{f}_2^L),$$ where $`\delta D_{V,A,VA}^{()}`$ and $`G_1^{()}`$ in $`c_2^\pm `$ are combinations of the SM and non-SM form factors in eq.(1) but without $`\delta D_v`$ while $`\xi ^{()}`$ is one including $`\delta D_v`$. This means $`\xi ^{()}`$ is a parameter to express $`CP`$ violation in the $`t\overline{t}v`$ couplings, and that is why the signs of the $`\xi ^{()}`$ terms for $`\mathrm{}^+`$ and $`\mathrm{}^{}`$ are opposite to each other. On the other hand, the coefficients $`a_{1,2,3}`$ consist of the SM parameters only, and $`f_{1,2,3}(x)`$ are analytic functions calculated in the SM. Similarly, the normalized double lepton-energy spectrum is given by the following formula: $$\frac{1}{\sigma }\frac{d^2\sigma }{dxd\overline{x}}=\underset{i=1}{\overset{6}{}}c_if_i(x,\overline{x}),$$ (8) where $`x`$ and $`\overline{x}`$ are for $`\mathrm{}^+`$ and $`\mathrm{}^{}`$ respectively, $$c_1=1,c_2=\xi ^{()},c_3=\frac{1}{2}\mathrm{Re}(f_2^R\overline{f}_2^L),$$ $$c_4=a_1^{}\delta D_V^{()}+a_2^{}\delta D_A^{()}+a_3^{}\mathrm{Re}(G_1^{()}),$$ $$c_5=a_1\delta D_V^{()}a_2[\delta D_A^{()}\mathrm{Re}(G_1^{()})]+a_3\mathrm{Re}(\delta D_{VA}^{()}),$$ $$c_6=\frac{1}{2}\mathrm{Re}(f_2^R+\overline{f}_2^L),$$ and again $`a_{1,2,3}^{}`$ are combinations of the SM parameters and $`f_i(x,\overline{x})`$ are analytic functions derived in the SM. ## 4. Parameter determination In order to study how precisely we can determine the coefficients $`c_i`$ in eqs.(7,8) when we have $`N`$ corresponding events, we used the optimal observable procedure . According to its prescription, we can deduce $`c_i`$ from the spectra (which we express as $`\mathrm{\Sigma }(\varphi )`$, where $`\varphi `$ means $`x`$ or $`(x,\overline{x})`$) with statistical uncertainty $$\mathrm{\Delta }c_i=\sqrt{X_{ii}/N},$$ where $`X_{ij}`$ is the inverse matrix of $$M_{ij}=𝑑\varphi \frac{f_i(\varphi )f_j(\varphi )}{\mathrm{\Sigma }(\varphi )}.$$ From the theoretical point of view, perfectly-polarized beams ($`P_{e^+}=P_e^{}=\pm 1`$) are the most attractive. However, those are difficult to achieve in practice, especially for the positron beam. So we discussed the following two cases: (1) $`P_{e^+}=0`$ vs $`P_e^{}=0,\pm 0.5,\pm 0.8`$ and $`\pm 1`$, (2) $`P_{e^+}=P_e^{}(P_e)=0,\pm 0.5,\pm 0.8`$ and $`\pm 1`$. In the analyses we assumed $`ϵ_{\mathrm{}}=0.6`$ as the lepton-tagging efficiency and $`L=100`$ fb<sup>-1</sup> as the integrated luminosity. Now let me show the main feature of the results focusing on the single spectrum: we found both $`\mathrm{\Delta }c_{2,3}^\pm `$ become smallest for $`P_e^{}=1`$/$`P_e=1`$. We can thereby conclude immediately that the best precision is obtained at $`c_3`$ measurements for these polarizations. However we have to be a bit more careful for $`c_2`$ measurements. This is because $`c_2`$ themselves vary depending on the polarization. Therefore we should discuss the statistical significance $`N_{SD}|c_2^\pm |/\mathrm{\Delta }c_2^\pm `$ inevitably instead of statistical errors only. For this purpose we considered the following two sets of the couplings as an example: * $`\mathrm{Re}(\delta A_{\gamma ,Z})=\mathrm{Re}(\delta B_{\gamma ,Z})=\mathrm{Re}(\delta C_{\gamma ,Z})=\mathrm{Re}(\delta D_{\gamma ,Z})=0.1`$, * $`\mathrm{Re}(\delta A_\gamma )=\mathrm{Re}(\delta B_\gamma )=\mathrm{Re}(\delta C_\gamma )=\mathrm{Re}(\delta D_\gamma )=0.1`$, $`\mathrm{Re}(\delta A_Z)=\mathrm{Re}(\delta B_Z)=\mathrm{Re}(\delta C_Z)=\mathrm{Re}(\delta D_Z)=0.1`$. As a result, we found that the use of negatively-polarized beam(s) is not always optimal: for the parameter set (a) a good precision in $`c_2^+`$ measurements is obtained when $`P_e<0`$, but even in this case the precision in $`c_2^{}`$ measurements becomes better for $`P_e>0`$ or even $`P_e=0`$. Moreover in case (b) both $`c_2^+`$ and $`c_2^{}`$ get the highest precision for $`P_e=+1`$. Therefore one should carefully adjust optimal polarization to test any given model. In any case one can conclude that (as far as the coefficient sets discussed here are concerned) appropriate beam polarization(s) provides measurements of $`c_{2,3}^\pm `$ at least at $`2\sigma `$ and $`3\sigma `$ level for $`P_{e^+}=0`$ and $`P_{e^+}0`$, respectively except for $`c_2^{}`$ in case (a), where $`|c_2^{}|`$ becomes tiny due to an accidental cancellation. We reached a similar conclusion for the double spectrum too: It depends on the structure of tested models what polarization(s) is best to study $`t\overline{t}v`$-couplings. I skip showing the details here, however, for want of space. ## 5. Summary Next-generation linear colliders of $`e^+e^{}`$, NLC, are expected to work as the cleanest facilities for studying top-quark interactions. There, we will be able to perform detailed tests of the top-quark couplings to the vector bosons and either confirm the SM simple generation-repetition pattern or discover some non-standard interactions. In this talk, I have shown main points of our latest model-independent analyses of the single- and the double-leptonic energy spectra for arbitrary longitudinal beam polarizations . We found ($`i`$) the use of longitudinal beams could be very effective in order to increase precision of the determination of non-SM couplings. However ($`ii`$) optimal polarization depends on the model of new physics under consideration. Therefore polarization of the initial beams should be carefully adjusted for each tested model.
no-problem/9908/cond-mat9908454.html
ar5iv
text
# References Vortex Clustering: The Origin of the Second Peak in the Magnetisation Loops of Type Two Superconductors \[Short title: Vortex Clustering\] D.K. Jackson<sup>a</sup>, M. Nicodemi<sup>a</sup>, G. K. Perkins<sup>b</sup>, N.A.Lindop<sup>c</sup>, Henrik Jeldtoft Jensen<sup>a</sup> <sup>1</sup><sup>1</sup>1Corresponding author, email: h.jensen@ic.ac.uk <sup>a</sup> Department of Mathematics, Queen’s gate 180, SW7 2BZ, United Kingdom, <sup>b</sup> Blackett Laboratory, Imperial College, Prince Consort Road, SW7 2BZ, United Kingdom, <sup>c</sup> Department of Chemistry, University of Edinburgh, Edinburgh EH9 3JJ, United Kingdom PACS: 74.60.Ge, 74.60.-w, 74.25.Ha Abstract We study vortex clustering in type II Superconductors. We demonstrate that the “second peak” observed in magnetisation loops may be a dynamical effect associated with a density driven instability of the vortex system. At the microscopic level the instability shows up as the clustering of individual vortices at (rare) preferential regions of the pinning potential. In the limit of quasi-static ramping the instability is related to a phase transition in the equilibrium vortex system. When the external magnetic field penetrates a type II superconductor, magnetic vortex lines appear inside the bulk of the sample. These vortex lines are repulsive and quantised each carrying a multiple $`q`$ of the magnetic flux quantum $`\varphi _0`$. The energy per unit length of a vortex line is proportional to $`q^2`$ and accordingly one typically expects to have configurations with well separated single quantised vortex lines . We show below, however, that dynamical vortex clustering may become very important: vortices form inhomogeneous spatial structures which become relevant for the evolution of the system. This clustering may for instance cause the “second peak” observed in magnetisation loops and strongly affects the structure of the vortex system, magnetic relaxation, and the distribution of the local magnetic induction as measured in $`\mu `$ spin-relaxation experiments. The understanding of structural properties of vortices, underlying the presence of the “second peak”, is one of the central issues of current research in superconductivity, and is related to fundamental aspects of vortex matter ranging from dynamical behaviour to phase transitions. (see eg. ). In a superconducting sample in the presence of an external magnetic field, vortex lines penetrating from the surface into the bulk may be trapped on pinning centres leading to a spatially inhomogeneous vortex distribution and to a net magnetisation of the sample. As the external field is increased the vortex lines are squeezed together and therefore interact more strongly. Such a strong interaction will counteract the pinning forces and thus one expects, as in fact is usually observed, the magnetisation to decrease with increasing external field once vortices have fully penetrated the sample. However, upon further increase of the external field the magnetisation is often observed to increase again . This behaviour leads to the so-called fishtail structure in the magnetisation data for YBCO or the arrowhead structure in equivalent data for BISCCO samples. A similar peak structure has also been observed in low temperature superconductors . The second peak is one of the most important unsolved problems in vortex physics. It has been seen as a signature of a phase transition in the vortex system and its explanation has often been attempted by focusing on collective aspects of pinning or relaxation effects related to the sweeping of the field . Here we present evidence that dynamical effects are of crucial importance for the second peak, nevertheless the mechanism inducing the peak is, in the quasi-static limit, related to a density driven phase transition in the vortex system. In Fig. 1 we show for reference an example of half magnetisation loops in YBCO for a set of different temperatures. One sees how the maximum of the second peak moves to higher magnetic fields as the temperature is decreased. This suggests that the mechanism behind the increase in the magnetisation occurring after the first peak must also be active at zero-temperature. We demonstrate below that the origin of the second peak may be related to the possible grouping of vortices at favourable regions in the random pinning potential, leading to large local fluctuations in the vortex density. This “clustering” can occur in superconductors for which the ratio, $`\kappa =\lambda /\xi `$, between the magnetic penetration depth $`\lambda `$ and the coherence length $`\xi `$ is not too small (from our Molecular Dynamics simulations we expect roughly $`\kappa >10`$). The clustering is a dynamical effect induced by the ramping of the external field, as is done in magnetisation experiments. As vortices entering (or leaving) the sample approach another vortex trapped at a position in the pinning potential three possibilities can occur. Firstly, the trapped vortex may be pushed ahead if the vortex-vortex repulsion is strong enough. Secondly, the approaching vortex may move around the trapped vortex. Or, if the trapped vortex is pinned by a force stronger than the maximum vortex-vortex repulsion, the approaching vortex may move into the favourable position in the pinning landscape in the immediate vicinity of the already trapped vortex. In this way the clustering may significantly enhance the effect of the rare strong pinning regions in an otherwise weak pinning background. For simplicity we here consider straight parallel flux lines at separation $`r`$ for which the interaction energy per unit length is $$U_{vv}(r)=\frac{\varphi _0^2}{2\pi \lambda ^2}\left[K_0(r/\lambda ^{})K_0(r/\xi ^{})\right],$$ (1) where $`\lambda ^{}=\lambda /\sqrt{1b}`$ and $`\xi ^{}=\xi /\sqrt{2(1b)}`$ are the effective field dependent London penetration depth and coherence length, respectively. $`b=B/B_{c_2}`$ is the reduced magnetic induction relative to the upper critical induction $`B_{c_2}`$, and $`K_0`$ is a modified Bessel function. Vortices can be brought close together because of the attractive second term representing the interaction between the vortex cores. The maximum repulsive force calculated from Eq.(1) is an upper limit for the repulsive force between two vortex lines. The relative tilting and wiggling of vortex lines will lead to a significant decrease in the repulsion (see Eg. ). Therefore, we expect the clustering effect to be at least as significant in three dimensions as we demonstrate here the effect to be in two dimensions. In fact clustering of vortices has been directly observed in electron microscopic imaging by Tonomura . In Fig. 2 we show the magnetisation obtained from zero temperature Molecular Dynamics (MD) simulations. The vortex interaction is given by Eq.(1) and it is cut off at half the system size. We use over damped dynamics in a square two dimensional system, with periodic boundary conditions, of sides $`100\xi `$. The external field is ramped by introducing vortices into a central strip with no pins ; the magnetisation plot is calculated by considering the average density of vortices and its gradient in the pinned region. Fig. 2 clearly shows that an upturn in the magnetisation occurs as the external field is increased above the penetration field. The increase in the magnetisation coincides with the appearance of vortices clustered within areas of order $`\xi ^2`$. It is important to mention that if the second term in Eq. 1 is left out, (see dashed line in Fig. 2.), clustering of vortices cannot occur and no significant upturn in the magnetisation is observed in the simulations. At the highest field densities the potential in Eq.(1) loses its validity. This regime is also difficult to handle numerically. Hence we do not study the full half loop in the MD simulations but use below a simplified lattice model to study the increasing as well as the decreasing leg of the magnetisation. Let us now describe how the clustering of vortices prevents the magnetisation, $`M`$, from decaying with increasing field. We recall that approximately $`Mj_c`$, where $`j_c`$ is the critical current density produced by the volume pinning force $`F_p=Bj_c`$. When a pinning centre becomes occupied by one or more trapped vortices the attractive short range pinning centre is effectively transformed into a longer range repulsive centre. Other vortices approaching this pin will feel a repulsive vortex-vortex force proportional to the number of trapped vortices at a distance $`\lambda `$ rather than the attractive force of range $`\xi `$ from the initially “empty” pinning centre. The local pinning strength will fluctuate through the sample with typically a high density of weak pinning centres and only a few sparse local strong regions. Vortices clustered at the few strong pinning regions can then form spatially extended energy barriers which cage other diffusing vortices. We emphasise that this picture is somewhat schematic and that in reality dynamical and collective effects are important. Vortices are moving in an ever changing energy landscape produced by the combined effect of the static spatial pinning potential and the instantaneous metastable configuration of the interacting vortices. The importance of dynamical effects follows from the fact that the detailed form, especially the width, of the magnetisation loops dependence on the ramping rate of the external field. This is the case in experiments (see ref.s in ) as well as in our simulations . To study in detail the change of the effective energetic panorama due to clustering and its consequences for the “second peak” we consider now a schematic model. This approach is close in spirit to similar lattice systems introduced to describe fluxons in superconductors, see e.g. . In particular we study an extension of a coarse grained cellular-automaton-like model recently introduced by Bassler and Paczuski (BP) . We consider a simplified version of a many body system with pair interactions given in eq. (1) representing a lattice model of repulsive particles in a pinning potential and in contact with a particle reservoir at a given density. Since our model explicitly allows multiple occupancy of lattice sites up to a value $`N_{c2}`$, we call it a Restricted Occupancy Model (ROM). We apply Monte Carlo dynamics and use a Hamiltonian of the form: $$=\frac{1}{2}\underset{ij}{}n_iA_{ij}n_j\frac{1}{2}\underset{i}{}A_{ii}n_i\underset{i}{}A_i^pn_i$$ (2) Here $`n_i\{0,\mathrm{},N_{c2}\}`$ is an integer occupancy variable equal to the number of particles on site $`i`$. The parameter $`N_{c2}`$ plays the role of $`B_{c2}`$, it bounds the particle density per site below a critical value (here $`N_{c2}=27`$). The first term in eq.(2) represents the repulsive vortex interaction energy. The second term in eq.(2) just normalises the particle self-interaction energy. Here we for simplicity choose the coarse graining length to be of order the zero temperature London penetration length $`\lambda (0)`$. This allows us to relate the restriction number $`N_{c2}`$ to the upper critical field $`B_{c2}`$ in the following way $`N_{c2}=B_{c2}\lambda (0)^2/\varphi _0`$ where $`\varphi _0=hc/2e`$ is the magnetic flux quantum. With this choice of coarse graining length it is natural as a first approximation to assume: $`A_{ii}=A_v=1`$, $`A_{ij}=A_n`$ if $`i`$ and $`j`$ are nearest neighbours and $`A_{ij}=0`$ for all other couples of sites. We will below briefly discuss the validity of this approximation of the $`A_{ij}`$ matrix. The third term in eq.(2) represents a random pinning potential acting on a fraction $`\rho =0.5`$ of the lattice sites with $`A^p=0.5`$ and $`A^p=0`$ elsewhere. (The same set of interactions is used in the BP model ). Two opposite sides of our square system ($`L=32^2`$) are in contact with a reservoir at a given density $`N_{ext}`$. Particles are introduced and escape the system through the reservoir only. Fig. 3 shows the results of our Monte Carlo simulations of this model. We ramp $`N_{ext}`$ and record the magnetisation, $`M=N_{in}N_{ext}`$ (with $`N_{in}=_in_i/L^2`$), as a function of $`N_{ext}`$. The ramping of the external reservoir density, $`N_{ext}`$, is simply done by increasing it from zero up to some given value (and then decreasing back) by a sequence of small increments, $`\mathrm{\Delta }N_0`$. After each increment the system is let to relax for a time $`\tau `$ (in unit of Monte Carlo sweeps). This corresponds to a sweep rate of the applied field of $`\gamma =\mathrm{\Delta }N_0/\tau `$. We recorded magnetisation loops for several values of the ratio $`\kappa ^{}(\mathrm{ln}A_v/A_n)^1`$ which controls the interaction potential. The parameter $`\kappa ^{}`$ is qualitatively similar to $`\kappa =\lambda /\xi `$. When $`\kappa ^{}`$ is large enough a definite second peak appears in $`M`$. These magnetisation loops looks qualitatively similar to the experimental loops, though the peak position and amplitude of peaks in Fig. 3 are much more asymmetric than the peaks of the experimental data in Fig. 1. Nevertheless, typically magnetisation loops in other experimental samples and compounds are clearly asymmetric, see e.g. . In fact even Fig. 1 exhibits, on careful inspection, slight asymmetries. The detailed loop shape depends not only on the choice of $`\kappa ^{}`$, as shown in Fig. 3, but also on the value of $`N_{c2}`$ and the ramping rate, $`\gamma `$. We find that in the $`\gamma 0`$ limit, the second peak location, is associated with a sharp jump in $`M_{eq}lim_{\gamma 0}M(\gamma )`$, corresponding to a true transition . The precise nature of this transition is currently under study. The transition occurs above the melting transition and leads to a significant increase in the effective energy barriers experienced by the diffusing vortices. Increasing $`N_{c2}`$ (corresponding to higher values of $`B_{c2}`$) increases the separation between the first and the second peak. The specific features of the loops do also depend on the ratio between the characteristic relaxation time of the vortex system and the ramping rate, for details see . Quantitative differences between the simulations and experiments are to be expected. One reason is that the interaction strength between vortex lines depends on the magnetic induction and the temperature. As for instance in the London approximation of Eq.(1) through the field and temperature dependence of $`\lambda ^{}`$ and $`\xi ^{}`$. This corresponds to a field and temperature dependence of the individual elements of the coupling matrix $`A_{ij}`$ in Eq.(2). Moreover, the effective vortex screening length, $`\lambda ^{}`$ (see Eq. 1), increases with increasing magnetic field. This effect implies that non-zero $`A_{ij}`$ elements between sites of separation larger than nearest neighbour may become relevant as the field is increased. Interestingly, however, the present simple approximation captures the qualitative features of the magnetic properties. We emphasise that according to the above picture the second peak is a dynamical effect associated with the vortices being forced in and out of the sample. It is linked to a density driven instability of the vortex system. At the more microscopic level of description used in the Molecular Dynamics simulations the instability shows up as the clustering of individual vortices at (rare) preferential regions of the pinning potential. In the coarse grained description of the considered lattice model the instability is related to an underlying density driven phase transition of the equilibrium system. The instability induces a dramatic change of the effective collective energy landscape encountered by the diffusing vortices. This, in turn, enormously enhances equilibrium times and induces the presence of significant more spatial disorder, in strict correspondence with glass formers . This clustering or density instability can produce the second peak in the magnetisation loops. It can be thought of as a type of strong plastic deformation, an effect which should be observable in neutron scattering or $`\mu `$-spin resonance experiments probing the distribution of local magnetic induction. DKJ and NL are grateful to acknowledge EPSRC and Imperial College Centre for High Temperature Superconductivity for studentships. HJJ is supported by EPSRC. Figures Captions Fig. 1 An example of magnetisation data and the fishtail effect in a single crystal of YBa2Cu4O8 at temperautres 20K 50K and 65K. The magnetic moment is measured continuously using a vibrating sample magnetometer while the applied field is ramped from zero to eight Tesla and back to zero at a rate 10mT/sec . The 124 system is neither susceptible to twinning or oxygen inhomogeneity (unlike YBa2Cu3O7-d) indicating that the fishtail effect is intrinsic to the pinning of point-like disorder in the crystal lattice. Fig. 2 Magnetisation ($`dB/dx`$) vs field ($`B`$) for simulation with $`\kappa `$ of $`67`$. The external field is ramped by adding one vortex between relaxation intervals of $`40`$ where each time step is a maximum of $`0.01`$. Fields calculated for $`\xi `$ of $`15`$ Å. Solid line represents the system interacting through Eq. 1 (soft core). The dashed line represents the system interacting through Eq. 1 with last term omitted (hard core). The short-long dashed curve represents the proportion of stacked vortices in the soft core case. Pinning centres are represented by Gaussian wells of width $`\xi `$ and of amplitude $`0.3\xi ^2`$ times the condensation energy. In this simulation they can exert a maximum pinning force (at zero external field) of $`410^{10}A/cm^2`$ for the $`\kappa =10`$ and $`810^8A/cm^2`$ for $`\kappa =100`$. Fig. 3 The magnetisation, $`M`$, as a function of the applied field density, $`N_{ext}`$, in the 2D R.O.M. model for $`\kappa ^{}=0.43,0.76,0.79`$. The ramp rate for $`N_{ext}`$ is $`\mathrm{\Delta }N_0/\tau =10^3`$.
no-problem/9908/quant-ph9908030.html
ar5iv
text
# Are violations to temporal Bell inequalities there when somebody looks? ## 1 Introduction The validity of quantum mechanics at the macroscopic level is still an open question. Leggett and Garg have challenged this question by proposing laboratory tests aimed at comparing, in a macroscopic domain, the predictions of a set of theories incorporating realism and non-invasivity, two properties not shared by quantum mechanics, versus quantum mechanics itself. The test involved certain inequalities, called temporal Bell inequalities in analogy to the well-known spatial ones . These last have been already experimentally studied to understand the validity of quantum mechanics against local realism at the microscopic level. Temporal Bell inequalities instead, involving correlation probabilities between subsequent measurements of the same observable, hold in realistic theories but are violated under certain conditions by quantum mechanics. The general ingredients required to discuss temporal Bell inequalities, regardless of the concrete scheme used, are three: (a) the possibility to have a coherent, quantum evolution of the state; (b) a dichotomic observable, already discrete or derived by a continuous one; (c) different-time correlation probabilities between subsequent measurements of the dichotomic observable. Leggett and Garg originally proposed use of an rf-SQUID, further analyzed in , while more recently other configurations have been discussed, such as a two-level atomic system undergoing optically driven Rabi oscillations , and Rydberg atoms interacting with a single quantized mode of a superconducting resonant cavity . The proposal of Leggett and Garg has been criticized focusing mainly on the assumptions made in the macrorealistic approaches. In this Letter we instead consider the full quantum mechanical predictions for the various proposed experiments. The main point of our analysis is that, as in any quantum mechanical prediction, probability distributions for the outcome of subsequent measurements have to be evaluated and compared to the analogous quantities actually measurable in laboratory. In practice, we will restrict ourselves to the evaluation of average and variance of the observed quantities. While the average value is necessary to define the violation itself, its variance is crucial to establish if the corresponding experimental resolution is sufficient to assign in an unambiguous way the value of the dichotomic observable . We will check if, under the experimental conditions which would lead to a contradiction between quantum and classical predictions for the correlation probabilities, the actual resolution can be sufficient to distinguish the two values of the dichotomic observable – otherwise it would be impossible to state that the quantum mechanical probabilities refer to well–defined observables. This feature was not taken into account in any of the previous papers of other authors on the subject , and constitutes the main point of our work. Indeed, we have already addressed this point elsewhere referring to specific physical schemes; the novelty of the present contribution is represented by the generality of our consideration, which apply to all experimental proposals put forward so far. ## 2 Quantum measurements on bistable systems We begin by briefly reviewing the conventional analysis of quantum measurements on a generic bistable system undergoing oscillations with period $`\tau `$ between the two states $`|+`$ and $`|`$, according to the time evolution operator $`\widehat{U}(t_\mathrm{b},t_\mathrm{a})`$. The effect of a measurement of the dichotomic observable $`\widehat{X}`$ is, as usual, represented through a non-unitary measuring operator $$\widehat{\mathrm{\Pi }}_X^{}=\widehat{\mathrm{\Pi }}_X,\widehat{\mathrm{\Pi }}_X^2=1\text{l},\widehat{\mathrm{\Pi }}_X|x=\delta _{Xx}|x,$$ (1) where $`|x=|\pm `$ and $`X`$ is one of the two eigenvalues $`\pm |X|`$. Given some initial preparation $`|X_\mathrm{a}`$ at time $`t_\mathrm{a}`$, after measurements performed at times $`t_\mathrm{b}`$, $`t_\mathrm{c}`$ with results $`X_\mathrm{b}`$, $`X_\mathrm{c}`$ respectively we get $$|x(t_\mathrm{b}^+)_{X_\mathrm{a}X_\mathrm{b}}=\widehat{\mathrm{\Pi }}_{X_\mathrm{b}}\widehat{U}(t_\mathrm{b},t_\mathrm{a})|X_\mathrm{a};|x(t_\mathrm{c}^+)_{X_\mathrm{a}X_\mathrm{b}X_\mathrm{c}}=\widehat{\mathrm{\Pi }}_{X_\mathrm{c}}\widehat{U}(t_\mathrm{c},t_\mathrm{b})|x(t_\mathrm{b}^+)_{X_\mathrm{a}X_\mathrm{b}}.$$ (2) The corresponding correlation probabilities are $$p_{X_\mathrm{a}X_\mathrm{b}}^{\mathrm{ab}}=x(t_\mathrm{b}^+)_{X_\mathrm{a}X_\mathrm{b}}^2;p_{X_\mathrm{a}X_\mathrm{b}X_\mathrm{c}}^{\mathrm{bc}}=x(t_\mathrm{c}^+)_{X_\mathrm{a}X_\mathrm{b}X_\mathrm{c}}^2,$$ (3) and their variances, whose square roots are called effective uncertainties, are $$\left(\mathrm{\Delta }X_{\mathrm{eff}}^{\mathrm{ab}}\right)^2=\underset{x}{}\left(X_\mathrm{b}x\right)^2p_{X_\mathrm{a}x}^{\mathrm{ab}};\left(\mathrm{\Delta }X_{\mathrm{eff}}^{\mathrm{bc}}\right)^2=\underset{x}{}\left(X_\mathrm{c}x\right)^2p_{X_\mathrm{a}X_\mathrm{b}x}^{\mathrm{bc}}.$$ (4) Their value is dictated by the laws of quantum mechanics, regardless of the size of the ensemble over which measurements are performed. From the definition Eq. (4) it follows that $`0\mathrm{\Delta }X_{\mathrm{eff}}2|X|`$ at any times. On the other hand, a resolution large enough is required in order to assign to the dichotomic observable a non-zero value without ambiguity. In other words, one has to take as non-ambiguous only measurements for which a condition $`\mathrm{\Delta }X_{\mathrm{eff}}\xi `$ is satisfied, the value of the threshold effective uncertainty $`\xi `$ defining the resolution criterion. A null value of $`\xi `$ means that only infinite-resolution measurements are taken as non-ambiguous; when $`\xi =2|X|`$ no discrimination is made between measurements with or without a resolution power sufficient to distinguish the two states $`|+`$ and $`|`$. Any of the criteria commonly used in resolution studies can be related to a value of the threshold between these two extremes. For example, the well-known half-width criterion corresponds to $`\xi =\xi _{1/2}(2\mathrm{ln}2)^{1/2}|X|`$: two probabilities distribution centered at $`\pm |X|`$ can be resolved if and only if their variance satisfies $`\mathrm{\Delta }X<\xi _{1/2}`$. A less stringent criterion would be for instance to require $`\mathrm{\Delta }X<|X|`$. ## 3 Temporal Bell inequalities Temporal Bell inequalities are formally derived by assuming the existence of $`n`$-times correlation probabilities $`p_{X_1\mathrm{}X_n}^{t_1\mathrm{}t_n}`$ satisfying positive definiteness and completeness, as well as the non-invasive measurability at any intermediate time . Under these conditions, for instance, $$p_+^{\mathrm{ac}}=p_+^{\mathrm{abc}}+p_{++}^{\mathrm{abc}}p_+^{\mathrm{abc}}+p_{++}^{\mathrm{abc}}+p_{++}^{\mathrm{abc}}+p_+^{\mathrm{abc}}=p_+^{\mathrm{ab}}+p_+^{\mathrm{bc}}$$ (5) which involves measurements performed at the subsequent times $`t_\mathrm{b}`$ and $`t_\mathrm{c}`$ on a system prepared in state $`|+`$ at the initial time $`t_\mathrm{a}`$. More in general three types of inequalities, analogous to (5), can be derived: $`\mathrm{\Delta }P_\mathrm{I}(t_{\mathrm{ab}},t_{\mathrm{bc}})`$ $``$ $`p_{X_\mathrm{a},X_\mathrm{c}}^{\mathrm{ac}}p_{X_\mathrm{a},X_\mathrm{b}}^{\mathrm{ab}}p_{X_\mathrm{b},X_\mathrm{c}}^{\mathrm{bc}}0;`$ $`\mathrm{\Delta }P_{\mathrm{II}}(t_{\mathrm{ab}},t_{\mathrm{bc}})`$ $``$ $`p_{X_\mathrm{a},X_\mathrm{b}}^{\mathrm{ab}}p_{X_\mathrm{a},X_\mathrm{c}}^{\mathrm{ac}}p_{X_\mathrm{b},X_\mathrm{c}}^{\mathrm{bc}}0;`$ (6) $`\mathrm{\Delta }P_{\mathrm{III}}(t_{\mathrm{ab}},t_{\mathrm{bc}})`$ $``$ $`p_{X_\mathrm{b},X_\mathrm{c}}^{\mathrm{bc}}p_{X_\mathrm{a},X_\mathrm{b}}^{\mathrm{ab}}p_{X_\mathrm{a},X_\mathrm{c}}^{\mathrm{ac}}0.`$ While inequalities of type III are ruled out by the experimental requirement that the system is prepared in a definite state at time $`t_\mathrm{a}`$, the other two types are in principle well suitable for experimental test. In fact, as originally observed in , each one of them is violated for some subset of measurement times $`t_{\mathrm{ab}}t_\mathrm{b}t_\mathrm{a}`$, $`t_{\mathrm{bc}}t_\mathrm{c}t_\mathrm{b}`$. On the other hand, the distinguishability criterion is fulfilled for another subset of $`(t_{\mathrm{ab}},t_{\mathrm{bc}})`$ depending upon the chosen threshold effective uncertainty $`\xi `$. It is therefore useful to introduce an overlap integral expressing the average probability difference $`\mathrm{\Delta }P_\alpha `$ (where $`\alpha =\mathrm{I},\mathrm{II},\mathrm{III}`$) integrated over all the time intervals for which both temporal Bell inequalities are violated by quantum mechanical predictions and distinguishability is assured by $`\mathrm{\Delta }X_{\mathrm{eff}}\xi `$: $$𝒪_\alpha (\xi )\frac{1}{\tau ^2}_{}\mathrm{\Delta }P_\alpha dt_{\mathrm{ab}}dt_{\mathrm{bc}}$$ (7) where the subscript $``$ means that the integration is restricted to the values of $`(t_{\mathrm{ab}},t_{\mathrm{bc}})`$ for which $`\mathrm{\Delta }P_\alpha >0`$ and $`\mathrm{\Delta }X_{\mathrm{eff}}\xi `$ for all measurements in $`\mathrm{\Delta }P_\alpha `$. The meaning of the overlap integral Eq. (7) is the following: given a threshold effective uncertainty $`\xi `$, it provides a measure of whether the violation of a certain temporal Bell inequality is compatible with the discrimination between the two values of the dichotomic observable, according to the criterion defined by $`\xi `$. From Eq. (7), it follows that $`𝒪_\alpha (\xi )`$ is continuous and monotonically increasing with $`\xi `$. Based on the considerations below Eq. (4), we expect also $`𝒪_\alpha (\xi 0)=0`$, $`𝒪_\alpha (\xi 2|X|)=\tau ^2_0^\tau dt_{\mathrm{ab}}_0^\tau dt_{\mathrm{bc}}\mathrm{\Delta }P_\alpha >0`$. Therefore the overlap integral will be nonzero only for $`\xi `$ large enough. We thus define $$\xi _\alpha \mathrm{min}\{\xi :𝒪_\alpha (\xi ^{})>0,\xi ^{}>\xi \}=\mathrm{max}\{\xi :𝒪_\alpha (\xi ^{})<0,\xi ^{}<\xi \}.$$ (8) This has to be compared with the $`\xi _{1/2}`$ defined above, in order to find out if violation of the inequalities Eq. (6) and discrimination of the two values of $`X`$ can be obtained in the same experiment. As a tool for our analysis we first discuss a toy model exhibiting all the relevant features intrinsic to the temporal Bell inequalities, moving later to the actual Hamiltonian of a rf-SQUID system as proposed in . We do not take into account finite temperature or imperfect efficiency effects, irrelevant to the arguments we plan to discuss since they can only diminish the chance to observe the predicted violations (see for instance the discussion in ). ## 4 Spin-1/2 particle case Let us consider a spin-1/2 particle precessing in a uniform magnetic field. Its Hamiltonian is $`\widehat{H}=\stackrel{}{\sigma }\stackrel{}{B}`$ with $`\stackrel{}{\sigma }=(\sigma _x,\sigma _y,\sigma _z)`$ the Pauli matrices and $`\stackrel{}{B}=(B,0,0)`$. The temporal evolution of the state vector is ruled by the operator $$\widehat{U}(t_\mathrm{b},t_\mathrm{a})=\mathrm{exp}\left[\frac{i}{\mathrm{}}B\sigma _x(t_\mathrm{b}t_\mathrm{a})\right].$$ (9) A measurement of the third component of the spin is represented by the projector $$\widehat{\mathrm{\Pi }}_\mathrm{\Sigma }=\frac{1+2\mathrm{\Sigma }\sigma _z}{2},\mathrm{\Sigma }=\pm \frac{1}{2}.$$ (10) The correlation probabilities appearing in (5) are then: $$p_+^{\mathrm{ab}}=\mathrm{sin}^2\mathrm{\Omega }t_{\mathrm{ab}};p_+^{\mathrm{ac}}=\mathrm{sin}^2\mathrm{\Omega }(t_{\mathrm{ab}}+t_{\mathrm{bc}});p_+^{\mathrm{bc}}=\mathrm{cos}^2\mathrm{\Omega }t_{\mathrm{ab}}\mathrm{sin}^2\mathrm{\Omega }t_{\mathrm{bc}}.$$ (11) where the angular frequency $`\mathrm{\Omega }B/\mathrm{}`$ has been introduced, whence the oscillation period $`\tau _\sigma =2\pi /\mathrm{\Omega }=2\pi \mathrm{}/B`$. The effective uncertainties are then calculated according to (4). As already pointed out in , this discussion applies to the proposals and , dealing with the two-level dynamics of atomic systems. ## 5 SQUID case In the rf-SQUID case instead the monitored observable is the magnetic flux $`\mathrm{\Phi }`$ threading the ring, a continuous observable. Its sign is a dichotomic variable which is directly measurable in a quantum coherence experimental setup ; the corresponding projector is $$\widehat{\mathrm{\Pi }}_\mathrm{\Phi }=\mathrm{\Theta }(\mathrm{\Phi }\widehat{\phi }),\mathrm{\Phi }=\pm 1.$$ (12) The effective Hamiltonian describing the system is $$\widehat{H}=\frac{\mathrm{}^2}{2C}\frac{\mathrm{d}^2}{\mathrm{d}\phi ^2}+\frac{(\phi \mathrm{\Phi }^{\mathrm{ext}})^2}{2L}\frac{I_\mathrm{c}\mathrm{\Phi }_0}{2\pi }\mathrm{cos}\left(2\pi \frac{\phi }{\mathrm{\Phi }_0}\right),$$ (13) where $`C`$ is the weak link capacitance (which plays the role of an effective mass), $`L`$ the ring inductance, $`I_\mathrm{c}`$ the critical current of the junction, $`\mathrm{\Phi }^{\mathrm{ext}}`$ is the external magnetic flux threading the ring and the flux quantum $`\mathrm{\Phi }_0=\mathrm{}/2e2.0710^{15}`$ Wb. A bistable regime is obtained if the condition $`1<2\pi LI_\mathrm{c}/\mathrm{\Phi }_0<5\pi /2`$ is fulfilled. In this case, by introducing the adimensional variable $`\varphi =(\phi \mathrm{\Phi }^{\mathrm{ext}})/\mathrm{\Phi }_0`$ and by fixing the value of the external flux at $`\mathrm{\Phi }^{\mathrm{ext}}/\mathrm{\Phi }_0=(n+1/2)`$, $`n`$ Z the potential can be approximated – up to a constant – by the quartic double-well form $$V(\varphi )=\frac{\pi ^3}{3}I_\mathrm{c}\mathrm{\Phi }_0(\varphi \varphi _0)^2(\varphi +\varphi _0)^2,$$ (14) where the two minima $`\pm \varphi _0`$ are the solutions of the equation $`\mathrm{sin}(2\pi \varphi )/\varphi =\mathrm{\Phi }_0/LI_\mathrm{c}`$ and are separated by a barrier of height $`\mathrm{\Delta }V=(\pi ^3/3)I_\mathrm{c}\mathrm{\Phi }_0\varphi _0^4`$. Correspondingly, the ground state has two peaks localized around $`\pm \varphi _0`$, each of width $`\sigma _0`$ determined by the approximate relationship $`C\omega _0^2\sigma _0^2/2\mathrm{}\omega _0/4`$ where $`\omega _0`$ is the plasma frequency, i.e. the angular frequency of the small oscillations around $`\pm \varphi _0`$. A flux wavepacket localized at $`\pm \varphi _0`$ is a superposition of the ground and first excited state and has width $`\sigma _0`$. If no measurement is performed, such a state will undergo coherent tunneling oscillations between the two wells with period $`\tau _\varphi 2\pi \mathrm{}/\mathrm{\Delta }E_0`$, where $`\mathrm{\Delta }E_0`$ is the splitting between the lowest two energy eigenvalues. This last decreases exponentially with increasing barrier height, whereas the tunneling frequency should be at least of the order of magnitude of 1 MHz – otherwise the coherent oscillations would be damped by the interaction with the environment. On the other hand, in order to treat the system as a two-level one , the probability for finding the system in the barrier region around $`\varphi =0`$ should be negligible; in other words, it should be $`\sigma _0^2\varphi _0^2`$, which in turn could be achieved with a high enough barrier. It is possible to choose the parameters in (13) in order to satisfy both these requirements: indeed, with $`L=150pH`$, $`C=0.15pF`$ and $`I_\mathrm{c}=2.5\mu A`$ (very close to the values used in ) it is $`\tau _\varphi ^15.94\text{MHz}`$ and $`\sigma _0^2/\varphi _0^20.08`$. ## 6 Results and discussion In fig. 1 the overlap integral defined in Eq. (7) is plotted as function of $`\xi `$ for all the inequalities of the form (5). The result depicted here is valid both for the spin toy-model and for the actual SQUID Hamiltonian (with the parameters quoted above), which lead to identical predictions because, as the tunneling oscillations dominate the dynamics, the two-level approximation is appropriate to describe the flux correlation probabilities in the rf-SQUID. No dependence on the actual values of $`X_\mathrm{a}`$, $`X_\mathrm{b}`$ and $`X_\mathrm{c}`$ is seen, but only on the type of inequality. The type I turns out to be more favourable for the observation of violations. Nevertheless, as it can be seen from the Figure, in both cases $`\xi _\alpha >1.4|X|`$, which is greater than any reasonable resolution threshold. This is the main result of this Letter showing that, even in the most favourable situation, it is impossible to observe violations to temporal Bell inequalities predicted by quantum mechanics maintaining at the same time the resolution high enough to distinguish the two values of the dichotomic observable. The correlation probabilities (3) are obtained by means of the standard rules of quantum mechanics from the corresponding interfering amplitudes. The violation of the inequality (5) however may be understood by noticing that it necessarily implies that at least one of the three time correlation probabilities $`p_{ijk}^{\mathrm{abc}}`$ becomes negative, namely becomes a Wigner function (pseudoprobability). This means, according to Feynman , that “the assumed condition of preparation or verification are experimentally unattainable” as a consequence of the uncertainty principle which forbids the existence of joint probabilities for incompatible variables. For instance, in our case from (5) follows that, when $`X_\mathrm{a}=X_\mathrm{b}=X_\mathrm{c}=|X|`$ in (6), $`\mathrm{\Delta }P_\mathrm{I}=p_{++}^{\mathrm{abc}}`$: therefore in the region where the temporal Bell inequality is violated, at least one of the joint pseudoprobabilities at three different times is negative. This result holds in general – and is no surprise. One should remember in fact that Wigner’s proof of validity of spatial Bell inequalities is based on the assumption, analogous to that made before Eq. (5), of non-negative joint probabilities for the spin components along different directions. Their violation should be ascribed, therefore, to the uncertainty principle which prevents the existence of these non-negative joint probabilities. This consequence of the uncertainty principle is independently confirmed by a recent result according to which the violation of spatial Bell inequalities is directly connected to the appearance of negative conditional entropies, a feature which is classically forbidden. On the other hand, and this is the difference between spatial and temporal Bell inequalities, it is the uncertainty principle itself which introduces in our case a constraint on the ability to resolve the two states in a second measurement at a later time on the same particle. (Such a constraint does not exist when the two measurements are performed at the same time on two different particles located in different space points). This second constraint arises from the fact that, as discussed in detail in , a quantum measurement of a given observable induces a perturbation in the evolution of its canonical conjugate, and this in turn produces an uncertainty on the outcome of a measurement of the same variable at a later time. This is crucial because, as we mentioned at the beginning of this paper, repeating measurements on the same observable is precisely what discriminates temporal Bell inequalities from the spatial version, where two different systems are observed only once. These two constraints, both arising from the uncertainty principle, are, as we have seen, conflicting, because the region in which the violations arise and the region in which the resolution is high enough to resolve between the two states do not overlap. These general considerations should affect other situations in which a macroscopic quantum system is repeatedly measured, for instance an atomic Bose-Einstein condensate in which the Josephson dynamics is studied. At the very core of quantum mechanics, the uncertainty principle has two competing consequences when applied to a single macroscopic system repeatedly monitored, the violation of classical probability laws for predictions on a dichotomic observable and the limitations on the resolution of the observable itself, and it seems impossible to experimentally unravel these two features without conflict. \*** We thank C. Cosmelli and C. Kuklewicz for useful discussions, and L. Chiatti for comparison of codes simulating the behaviour of the rf-SQUID. One of us (T. C.) thanks the Institute of Nuclear Theory at the University of Washington and the Department of Physics at the University of Trento (in particular M. Traini and S. Stringari) for their hospitality, the Department of Energy and the ECT\* and INFN in Trento for partial support during the completion of this work, and A. J. Leggett and A. Garg for stimulating discussions.
no-problem/9908/hep-ph9908471.html
ar5iv
text
# Untitled Document ## Abstract The fermion mass problem and the ideas of mass protection are briefly reviewed. The Fritzsch ansatz for the quark mass matrices and a recent variant, based on a lightest flavour mixing mechanism in which all the CKM mixing angles disappear in the chiral symmetry limit of vanishing up and down quark masses, are discussed. The Anti-Grand Unification Model (AGUT) and the Multiple Point Principle (MPP) used to calculate the values of the Standard Model gauge coupling constants in the theory are described. The application of the MPP to the pure Standard Model predicts the top quark mass to be $`173\pm 5`$ GeV and the Higgs particle mass to be $`135\pm 9`$ GeV. Mass protection by the chiral quantum numbers of the maximal AGUT gauge group $`SMG\times U(1)_f`$ provides a successful fit to the charged fermion mass spectrum, with an appropriate choice of Higgs fields to break the AGUT gauge group down to the Standard Model gauge group (SMG) close to the Planck scale. The puzzle of the neutrino masses and mixing angles presents a challenge to the AGUT model and approaches to this problem are briefly discussed. Published in the Proceedings of the Corfu Summer Institute on Elementary Particle Physics, 1998, J. High Energy Phys. Conf. Proc. corfu98/032. ## 1 Introduction As I discussed in my talk at the previous Corfu workshop in 1995, the pattern of observed quark and lepton masses, their mixing and three generation structure form the major outstanding problem of particle physics. The hierarchical structure of the charged fermion masses, ranging over five orders of magnitude from 1/2 MeV for the electron to 175 GeV for the top quark, and of the quark weak coupling matrix elements strongly suggests the existence of physics beyond the Standard Model (SM). Furthermore the growing experimental support for the existence of neutrino oscillations and hence for a non-zero neutrino mass, from SuperKamiokande and other data, provides direct evidence for non-Standard Model physics. So the experimental values of the SM fermion masses and mixing angles presently provide our best clues to the fundamental physics of flavour. A fermion mass term $$_{mass}=m\overline{\psi }_L\psi _R+h.c.$$ (1) couples together a left-handed Weyl field $`\psi _L`$ and a right-handed Weyl field $`\psi _R`$, which then satisfy the Dirac equation $$i\gamma ^\mu _\mu \psi _L=m\psi _R$$ (2) If the two Weyl fields are not charge conjugates $`\psi _L(\psi _R)^c`$ we have a Dirac mass term and the two fields $`\psi _L`$ and $`\psi _R`$ together correspond to a Dirac spinor. However if the two Weyl fields are charge conjugates $`\psi _L=(\psi _R)^c`$ we have a Majorana mass term and the corresponding four component Majorana spinor has only two degrees of freedom. Particles carrying an exactly conserved charge $`Q`$, like the electron, must be distinct from their anti-particles and can only have Dirac masses with $`\psi _L`$ and $`\psi _R`$ having equal charges $`Q_L=Q_R`$. However a neutrino could be a massive Majorana particle. The left-handed and right-handed top quark, $`t_L`$ and $`t_R`$ carry unequal SM $`SU(2)\times U(1)`$ gauge charges: $$Q_LQ_R(\mathrm{Chiral}\mathrm{charges})$$ (3) Electroweak gauge invariance protects the quarks and leptons from gaining a fundamental mass term ($`\overline{t_L}t_R`$ is not gauge invariant). This mass protection mechanism is of course broken by the Higgs effect, which naturally generates a mass for the top quark of the same order of magnitude as the SM Higgs field vacuum expectation value (vev). Thus the Higgs mechanism explains why the top quark mass is suppressed, relative to the fundamental (Planck, GUT…) mass scale of the physics beyond the SM, down to the scale of electroweak gauge symmetry breaking. However the further suppression of the other quark-lepton masses remains a mystery, which it is natural to attribute to mass protection by another approximately conserved (gauge) charge (or charges) beyond the SM, as discussed in section 3. In this talk I will appeal to the gauge charges of the Anti-Grand Unification Theory (AGUT) for this mass protection. The AGUT model and its connection with the Multiple Point Principle (MPP) is discussed in section 4. The MPP predictions for the top quark and Higgs particle masses within the pure SM are then discussed in section 5. The Higgs field sector required to break the AGUT gauge group down to that of the SM is described in section 6. The structure of the quark and charged lepton mass matrices resulting from AGUT mass protection is presented in section 7. I will then consider the neutrino mass problem in section 8 and conclude in section 9. However let me begin, in the following section 2, by considering the structure of the fermion mass matrices and some of the ansätze suggested by phenomenology. ## 2 Mass matrix texture The hierarchical structure of the Standard Model fermion mass spectrum naturally suggests that the fermion mass matrix elements have a similar hierarchical structure, each typically having a different order of magnitude. The smaller elements may then contribute so weakly to the physical masses and mixing angles that they can effectively be neglected and replaced by zero—texture zeros. The best known ansatz incorporating such a texture zero is the two generation Fritzsch hermitean ansatz : $$M_U=\left(\begin{array}{cc}0& B\\ B^{}& A\end{array}\right)M_D=\left(\begin{array}{cc}0& B^{}\\ B^{}& A^{}\end{array}\right)$$ (4) The assumed hierarchical structure gives the following conditions: $$|A||B|,|A^{}||B^{}|$$ (5) among the parameters. It follows that the two generation Cabibbo mixing is given by the well-known Fritzsch formula $$\left|V_{us}\right|\left|\sqrt{\frac{m_d}{m_s}}e^{i\varphi }\sqrt{\frac{m_u}{m_c}}\right|$$ (6) where $`\varphi =\mathrm{arg}B^{}\mathrm{arg}B`$. This relationship fits the experimental value well, provided that the phase $`\varphi `$ is close to $`\frac{\pi }{2}`$. The generalisation of the Fritzsch ansatz to three generations: $$M_U=\left(\begin{array}{ccc}0& C& 0\\ C^{}& 0& B\\ 0& B^{}& A\end{array}\right)$$ (7) $$M_D=\left(\begin{array}{ccc}0& C^{}& 0\\ C^{}& 0& B^{}\\ 0& B^{}& A^{}\end{array}\right)$$ (8) with the assumed hierarchy of parameters: $$|A||B||C|,|A^{}||B^{}||C^{}|$$ (9) however leads to an additional relationship $$|V_{cb}|\left|\sqrt{\frac{m_s}{m_b}}e^{i\varphi _2}\sqrt{\frac{m_c}{m_t}}\right|$$ (10) which is excluded by the data for any value of the phase $`\varphi _2`$. Consistency with experiment can, for example, be restored by introducing a non-zero 2-2 mass matrix element . There are several ansätze, with texture zeros , which give testable relations between the masses and mixing angles . Here I will discuss a recent suggestion , which predicts all the CKM mixing matrix elements in terms of quark masses. It is a common belief, due to the success of eq. (6), that the smallness of the Cabibbo mixing matrix element $`V_{us}`$ is due to the lightness of the $`u`$ and $`d`$ quarks. However not only the 1-3 generation mixing $`V_{ub}`$ but also the 2-3 generation mixing $`V_{cb}`$ happen to be small compared to $`V_{us}`$. This led us to the idea that all the other mixings, and primarily the 2-3 mixing, could also be controlled by the up and down quark masses $`m_u`$ and $`m_d`$ and vanishes in the chiral symmetry limit $`m_u=m_d=0`$. Therefore we consider an ansatz in which the diagonal mass matrix elements for the second and third generations are practically the same in the gauge (unrotated) and physical bases. We propose that the three mass matrices for the Dirac fermions—the up quarks ($`U`$ = $`u`$, $`c`$, $`t`$), the down quarks ($`D`$ = $`d`$, $`s`$, $`b`$) and charged leptons ($`E`$ = $`e`$, $`\mu `$, $`\tau `$)—are each hermitian with three texture zeros of the following form: $$M_i=\left(\begin{array}{ccc}0& a_i& 0\\ a_i^{}& A_i& b_i\\ 0& b_i^{}& B_i\end{array}\right)i=U,D,E$$ (11) with the hierarchy $`B_iA_i\left|b_i\right|\left|a_i\right|`$ between the elements. Our ansatz requires the diagonal elements ($`A_i`$, $`B_i`$), of the mass matrices $`M_i`$, to be proportional to the modulus square of the off-diagonal elements ($`a_i`$, $`b_i`$): $$\frac{A_i}{B_i}=\left|\frac{a_i}{b_i}\right|^2i=U,D,E$$ (12) It follows that the Cabibbo mixing is given by the Fritzsch formula eq. (6) which fits the experimental value well, provided that the CP violating phase $`\varphi `$ is required to be close to $`\frac{\pi }{2}`$. Our most interesting prediction (with the mass ratios calculated at the electroweak scale ) is: $`\left|V_{cb}\right|`$ $``$ $`\left|\sqrt{{\displaystyle \frac{m_d}{m_b}}}e^{i\gamma }\sqrt{{\displaystyle \frac{m_u}{m_t}}}\right|`$ (13) $``$ $`\sqrt{{\displaystyle \frac{m_d}{m_b}}}=0.038\pm 0.007`$ in good agreement with the current data $`\left|V_{cb}\right|=0.039\pm 0.003`$ . If we also take the phase $`\gamma =\mathrm{arg}b_D\mathrm{arg}b_U`$ to be $`\frac{\pi }{2}`$, the uncertainty in our prediction of eq. (13) is reduced from 0.007 to 0.004. Another prediction for the ratio: $$\left|\frac{V_{ub}}{V_{cb}}\right|=\sqrt{\frac{m_u}{m_c}}$$ (14) is quite general for models with nearest-neighbour mixing. An alternative scenario, in which the hermitian mass matrix for the up quarks is changed to be of the form: $$M_U=\left(\begin{array}{ccc}0& 0& c_U\\ 0& A_U& 0\\ c_U^{}& 0& B_U\end{array}\right)$$ (15) leads to mixing angles given by the simple and compact formulae: $$\left|V_{us}\right|\sqrt{\frac{m_d}{m_s}}\left|V_{cb}\right|\sqrt{\frac{m_d}{m_b}}\left|V_{ub}\right|\sqrt{\frac{m_u}{m_t}}$$ (16) While the values of $`\left|V_{us}\right|`$ and $`\left|V_{cb}\right|`$ are practically the same as in our first scenario and in good agreement with experiment, a new prediction for $`\left|V_{ub}\right|`$ (not depending on the value of the CP violating phase) should allow experiment to differentiate between the two scenarios in the near future. ## 3 Mass matrix texture from chiral flavour charges As we pointed out in section 1, a natural resolution to the charged fermion mass problem is to postulate the existence of some approximately conserved chiral charges beyond the SM. These charges, which we assume to be the gauge quantum numbers in the fundamental theory beyond the SM, provide selection rules forbidding the transitions between the various left-handed and right-handed quark-lepton states, except for the top quark. In order to generate mass terms for the other fermion states, we have to introduce new Higgs fields, which break the fundamental gauge symmetry group $`G`$ down to the SM group. We also need suitable intermediate fermion states to mediate the forbidden transitions, which we take to be vector-like Dirac fermions with a mass of order the fundamental scale $`M_F`$ of the theory. In this way effective SM Yukawa coupling constants are generated, which are suppressed by the appropriate product of Higgs field vacuum expectation values measured in units of $`M_F`$. Consider, for example, the model obtained by extending the Standard Model gauge group $`SMG=SU(3)\times SU(2)\times U(1)`$ with a gauged abelian flavour group $`U(1)_f`$. This $`SMG\times U(1)_f`$ gauge group is broken to SMG by the vev of a scalar field $`\varphi _S`$ where $`\varphi _S<M_F`$ and $`\varphi _S`$ carries $`U(1)_f`$ charge $`Q_f(\varphi _S)`$ = 1. Suppose further that the $`U(1)_f`$ charges of the Weinberg Salam Higgs field and the left- and right-handed bottom quark fields are: $$Q_f(\varphi _{WS})=0Q_f(b_L)=0Q_f(b_R)=2$$ (17) Then it is natural to expect the generation of a mass for the $`b`$ quark of order: $$\left(\frac{\varphi _S}{M_F}\right)^2\varphi _{WS}$$ (18) via a tree level diagram involving the exchange of two $`\varphi _S`$ tadpoles, in addition to the usual $`\varphi _{WS}`$ tadpole, with two appropriately charged vector-like fermion intermediate states of mass $`M_F`$. We identify $`ϵ_f=\varphi _S/M_F`$ as the $`U(1)_f`$ flavour symmetry breaking parameter. In general we expect mass matrix elements of the form: $$M(i,j)=\gamma _{ij}ϵ_f^{n_{ij}}\varphi _{WS}$$ (19) between the $`i`$th left-handed and $`j`$th right-handed fermion components, where $$\gamma _{ij}=𝒪(1),n_{ij}=Q_f(\psi _{L_i})Q_f(\psi _{R_j})$$ (20) So the effective SM Yukawa couplings of the quarks and leptons to the Weinberg-Salam Higgs field $`y_{ij}=\gamma _{ij}ϵ_f^{n_{ij}}`$ can consequently be small even though all fundamental Yukawa couplings of the “true” underlying theory are of $`𝒪`$(1). However it appears not possible to explain the fermion mass spectrum with an anomaly free set of flavour charges in an $`SMG\times U(1)_f`$ model with a single Higgs field $`\varphi _S`$ breaking the $`U(1)_f`$ gauge symmetry. In fact it is possible to produce a realistic quark-lepton spectrum, but at the expense of introducing three Higgs fields with relatively prime $`U(1)_f`$ charges and most of the SM fermions carrying exceptionally large $`U(1)_f`$ charges. Another possibility is to introduce SMG-singlet fermions with non-zero values of the $`U(1)_f`$ charge to cancel the $`U(1)_f^3`$ gauge anomaly (as in $`MSSM\times U(1)_f`$ models , which also use anomaly cancellation via the Green-Schwarz mechanism ). However we shall consider the alternative of extending the SM gauge group further—in fact to that of the anti-grand unification model introduced in the next section. We shall take the point of view that, in the fundamental theory beyond the SM, the Yukawa couplings allowed by gauge invariance are all of order unity and, similarly, all the mass terms allowed by gauge invariance are of order the fundamental mass scale of the theory—say the Planck scale. Then, apart from the element responsible for the top quark mass, the quark-lepton mass matrix elements are only non-zero due to the presence of other Higgs fields having vevs smaller (typically by one order of magnitude) than the fundamental scale. These Higgs fields will, of course, be responsible for breaking the fundamental gauge group $`G`$—whatever it may be—down to the SM group. In order to generate a particular effective SM Yukawa coupling matrix element, it is necessary to break the symmetry group $`G`$ by a combination of Higgs fields with the appropriate quantum number combination $`\mathrm{\Delta }\stackrel{}{Q}`$. When this “$`\mathrm{\Delta }\stackrel{}{Q}`$” is different for two matrix elements they will typically deviate by a large factor. If we want to explain the observed spectrum of quarks and leptons in this way, it is clear that we need charges which—possibly in a complicated way—separate the generations and, at least for $`tb`$ and $`cs`$, also quarks in the same generation. Just using the usual simple $`SU(5)`$ GUT charges does not help because both ($`\mu _R`$ and $`e_R`$) and ($`\mu _L`$ and $`e_L`$) have the same $`SU(5)`$ quantum numbers. So we prefer to keep each SM irreducible representation in a separate irreducible representation of $`G`$ and introduce extra gauge quantum numbers distinguishing the generations, by adding extra Cartesian-product factors to the SM gauge group. ## 4 Anti-Grand unification model In the AGUT model the SM gauge group is extended in much the same way as Grand Unified $`SU(5)`$ is often assumed; it is just that we assume another non-simple gauge group $`G=SMG^3\times U(1)_f`$, where $`SMGSU(3)\times SU(2)\times U(1)`$, becomes active near the Planck scale $`M_{Planck}10^{19}`$ GeV. So we have a pure SM desert, without any supersymmetry, up to an order of magnitude or so below $`M_{Planck}`$. The existence of the $`SMG^3\times U(1)_f`$ group means that, near the Planck scale, each of the three quark-lepton generations has got its own gauge group and associated gauge particles with the same structure as the SM gauge group. There is also an extra abelian $`U(1)_f`$ gauge boson, giving altogether $`3\times 8=24`$ gluons, $`3\times 3=9`$ $`W`$’s and $`3\times 1+1=4`$ abelian gauge bosons. The couplings of the $`i`$’th proto-generation to the $`SMG_i=SU(3)_i\times SU(2)_i\times U(1)_i`$ group are identical to those to the SM group. Consequently we have a charge quantization rule, analogous to the SM charge quantisation rule (see eq. (22) below), for each of the three proto-generation weak hypercharge quantum numbers $`y_i`$. For the colourless particles we have the Millikan charge quantization of all charges being integer when measured in units of the elementary charge unit, but for coloured particles the charges deviate from being integer by $`1/3`$ of the elementary charge for quarks and by $`+1/3`$ for antiquarks. This rule can be expressed by introducing the concept of triality $`t`$, which characterizes the representation of the centre of the colour $`SU(3)`$ group, and is defined so that $`t=0`$ for the trivial representation or for decuplets, octets and so on, while $`t=1`$ for triplet ($`\underset{¯}{3}`$) or anti-sextet etc. and $`t=1`$ for anti-triplet ($`\overline{\underset{¯}{3}}`$) or sextet etc. Then the rule can be written in the form $$Q+t/3=0(\text{mod}1)$$ (21) where $`Q`$ is the electric charge $`Q=y/2+t_3/2`$ ($`t_3`$ is the third component of the weak isospin, SU(2), and y is the weak hypercharge). So we may write this SM charge quantization rule as $$y/2+d/2+t/3=0(\text{mod}1)$$ (22) where we have introduced the duality $`d`$, which is defined to be $`0`$ when the weak isospin is integer and $`d=1`$ when it is half integer. At first sight, this $`SMG^3\times U(1)_f`$ group with its 37 generators seems to be just one among many possible SM gauge group extensions. However, it is actually not such an arbitrary choice, as it can be uniquely specified by postulating 4 reasonable requirements on the gauge group $`GSMG`$. As a zeroth postulate, of course, we require that the gauge group extension must contain the Standard Model group as a subgroup $`GSMG`$. In addition it should obey the following 4 postulates: The first two are also valid for $`SU(5)`$ GUT: 1. $`G`$ should transform the presently known (left-handed, say) Weyl particles into each other. Here we take the point of view that we do not look for the whole gauge group $`G`$, say, but only for that factor group $`G^{}`$ = $`G/H`$ which transforms the already known quark and lepton Weyl fields in a nontrivial way. That is to say, we ask for the group obtained by dividing out the subgroup $`HG`$ which leaves the quark and lepton fields unchanged. This factor group $`G^{}`$ can then be identified with its representation of the Standard Model fermions, i.e. as a subgroup of the $`U(45)`$ group of all possible unitary transformations of the 45 Weyl fields for the Standard Model. If one took $`G`$ to be one of the extensions of SU(5), such as SO(10) or the E-groups as promising unification groups, the factor group $`G/H`$ would be SU(5) only; the extension parts can be said to only transform particles that are not in the Standard Model (and thus could be pure fantasy a priori). 2. No anomalies, neither gauge nor mixed. We assume that only straightforward anomaly cancellation takes place and, as in the SM itself, do not allow for a Green-Schwarz type anomaly cancellation . But the next two are rather just opposite to the properties of the $`SU(5)`$ GUT, thus justifying the name Anti-GUT: 3. The various irreducible representations of Weyl fields for the SM group remain irreducible under $`G`$. This is the most arbitrary of our assumptions about $`G`$. It is motivated by the observation that combining SM irreducible representations into larger unified representations introduces symmetry relations between Yukawa coupling constants, whereas the particle spectrum does not exhibit any exact degeneracies (except possibly for the case $`m_b=m_\tau `$). In fact AGUT only gets the naive $`SU(5)`$ mass predictions as order of magnitude relations: $`m_bm_\tau `$, $`m_sm_\mu `$, $`m_dm_e`$. 4. $`G`$ is the maximal group satisfying the other 3 postulates. With these four postulates a somewhat complicated calculation shows that, modulo permutations of the various irreducible representations in the Standard Model fermion system, we are led to our gauge group $`SMG^3\times U(1)_f`$. Furthermore it shows that the SM group is embedded as the diagonal subgroup of $`SMG^3`$, as required in our AGUT model. The AGUT group breaks down an order of magnitude or so below the Planck scale to the SM group. The anomaly cancellation constraints are so tight that, apart from various permutations of the particle names, the $`U(1)_f`$ charge assignments are uniquely determined up to an overall normalisation and sign convention. In fact the $`U(1)_f`$ group does not couple to the left-handed particles or any first generation particles, and the $`U(1)_f`$ quantum numbers can be chosen as follows: $$Q_f(\tau _R)=Q_f(b_R)=Q_f(c_R)=1$$ (23) $$Q_f(\mu _R)=Q_f(s_R)=Q_f(t_R)=1$$ (24) The AGUT group breaks down an order of magnitude or so below the Planck scale to the diagonal subgroup of the $`SMG^3`$ subgroup (the diagonal subgroup is isomorphic to the usual SM group). For this breaking we shall use a relatively complicated system of Higgs fields with names $`W`$, $`T`$, $`\xi `$, and $`S`$. In order to fit neutrino masses as well, we need an even more complicated system. It should however be said that, although at the very high energies just under the Planck energy each generation has its own gluons, own W’s etc., the breaking makes only one linear combination of a certain colour combination of gluons “survive” down to low energies. So below circa 1/10 of the Planck scale, it is only these linear combinations that are present and thus the couplings of the gauge particles—at low energy only corresponding to these combinations—are the same for all three generations. You can also say that the phenomenological gluon is a linear combination with amplitude $`1/\sqrt{3}`$ for each of the AGUT-gluons of the same colour combination. That then also explains why the coupling constant for the phenomenological gluon couples with a strength that is $`\sqrt{3}`$ times smaller than for the AGUT-gluons (see eq. (25) below) if, as we effectively assume, the three AGUT $`SU(3)`$ couplings were equal to each other. The SM gauge coupling constants do not, of course, unify, because we have not combined the groups U(1), SU(2) and SU(3) together into a simple group, but their values have been successfully calculated using the Multiple Point Principle . According to the MPP, the coupling constants should be fixed such as to ensure the existence of many vacuum states with the same energy density; in the Euclideanised version of the theory, there is a corresponding phase transition. So if several vacua are degenerate, there is a multiple point. The couplings at the multiple points have been calculated in lattice gauge theory for the groups $`SU(3)`$, $`SU(2)`$ and $`U(1)`$ separately. We imagine that the lattice has a truly physical significance in providing a cut-off for our model at the Planck scale. The SM fine structure constants correspond to those of the diagonal subgroup of the $`SMG^3`$ group and, for the non-abelian groups, this gives: $$\alpha _i(M_{Planck})=\frac{\alpha _i^{MultiplePoint}}{3}i=2,3$$ (25) The situation is more complicated for the abelian groups, because it is possible to have gauge invariant cross-terms between the different $`U(1)`$ groups in the Lagrangian density such as: $$\frac{1}{4g^2}F_{\mu \nu }^{gen1}(x)F_{gen2}^{\mu \nu }(x)$$ (26) So, in first approximation, for the SM $`U(1)`$ fine structure constant we get: $$\alpha _1(M_{Planck})=\frac{\alpha _1^{MultiplePoint}}{6}$$ (27) The agreement of these AGUT predictions with the data is shown in figure 1. ## 5 The MPP Prediction for the Top Quark and Higgs masses in the Standard Model The application of the MPP to the pure Standard Model , with a cut-off close to $`M_{Planck}`$, implies that the SM parameters should be adjusted, such that there exists another vacuum state degenerate in energy density with the vacuum in which we live. This means that the effective SM Higgs potential $`V_{eff}(|\varphi |)`$ should, have a second minimum degenerate with the well-known first minimum at the electroweak scale $`|\varphi _{vac\mathrm{\hspace{0.33em}1}}|=246`$ GeV. Thus we predict that our vacuum is barely stable and we just lie on the vacuum stability curve in the top quark, Higgs particle (pole) mass ($`M_t`$, $`M_H`$) plane. Furthermore we expect the second minimum to be within an order of magnitude or so of the fundamental scale, i.e. $`|\varphi _{vac\mathrm{\hspace{0.33em}2}}|M_{Planck}`$. In this way, we essentially select a particular point on the SM vacuum stability curve and hence the MPP condition predicts precise values for $`M_t`$ and $`M_H`$. For the purposes of our discussion it is sufficient to consider the renormalisation group improved tree level effective potential $`V_{eff}(\varphi )`$. We are interested in values of the Higgs field of the order $`|\varphi _{vac\mathrm{\hspace{0.33em}2}}|M_{Planck}`$, which is very large compared to the electroweak scale, and for which the quartic term strongly dominates the $`\varphi ^2`$ term; so to a very good approximation we have: $$V_{eff}(\varphi )\frac{1}{8}\lambda (\mu =|\varphi |)|\varphi |^4$$ (28) The running Higgs self-coupling constant $`\lambda (\mu )`$ and the top quark running Yukawa coupling constant $`g_t(\mu )`$ are readily computed by means of the renormalisation group equations, which are in practice solved numerically, using the second order expressions for the beta functions. The vacuum degeneracy condition is imposed by requiring: $$V_{eff}(\varphi _{vac\mathrm{\hspace{0.33em}1}})=V_{eff}(\varphi _{vac\mathrm{\hspace{0.33em}2}})$$ (29) Now the energy density in vacuum 1 is exceedingly small compared to $`\varphi _{vac\mathrm{\hspace{0.33em}2}}^4M_{Planck}^4`$. So we basically get the degeneracy condition, eq. (29), to mean that the coefficient $`\lambda (\varphi _{vac\mathrm{\hspace{0.33em}2}})`$ of $`\varphi _{vac\mathrm{\hspace{0.33em}2}}^4`$ must be zero with high accuracy. At the same $`\varphi `$-value the derivative of the effective potential $`V_{eff}(\varphi )`$ should be zero, because it has a minimum there. Thus at the second minimum of the effective potential the beta function $`\beta _\lambda `$ also vanishes: $$\beta _\lambda (\mu =\varphi _{vac\mathrm{\hspace{0.33em}2}})=\lambda (\varphi _{vac\mathrm{\hspace{0.33em}2}})=0$$ (30) which gives to leading order the relationship: $$\frac{9}{4}g_2^4+\frac{3}{2}g_2^2g_1^2+\frac{3}{4}g_1^412g_t^4=0$$ (31) between the top quark Yukawa coupling and the electroweak gauge coupling constants $`g_1(\mu )`$ and $`g_2(\mu )`$ at the scale $`\mu =\varphi _{vac\mathrm{\hspace{0.33em}2}}M_{Planck}`$. We use the renormalisation group equations to relate the couplings at the Planck scale to their values at the electroweak scale. Figures 2 and 3 show the running coupling constants $`\lambda (\varphi )`$ and $`g_t(\varphi )`$ as functions of $`\mathrm{log}(\varphi )`$. Their values at the electroweak scale give our predicted combination of pole masses : $$M_t=173\pm 5\text{GeV}M_H=135\pm 9\text{GeV}$$ (32) ## 6 AGUT gauge symmetry breaking by Higgs fields There are obviously many different ways to break down the large group $`SMG\times U(1)_f`$ to the much smaller SMG. However, we can greatly simplify the situation by assuming that, like the quark and lepton fields, the Higgs fields belong to singlet or fundamental representations of all non-abelian groups. The non-abelian representations are then determined from the $`U(1)_i`$ weak hypercharge quantum numbers, by imposing the charge quantization rule eq. (22) for each of the $`SMG_i`$ groups. So now the four abelian charges, which we express in the form of a charge vector $$\stackrel{}{Q}=(\frac{y_1}{2},\frac{y_2}{2},\frac{y_3}{2},Q_f)$$ can be used to specify the complete representation of $`G`$. The constraint that we must eventually recover the SM group as the diagonal subgroup of the $`SMG_i`$ groups is equivalent to the constraint that all the Higgs fields (except for the Weinberg-Salam Higgs field which of course finally breaks the SMG) should have charges $`y_i`$ satisfying: $$y=y_1+y_2+y_3=0$$ (33) in order that their SM weak hypercharge $`y`$ be zero. We wish to choose the quantum numbers of the Weinberg-Salam (WS) Higgs field $`\varphi _{WS}`$ so that it matches the difference in charges between the left-handed and right-handed physical top quarks. This will ensure that the top quark mass in the SM is not suppressed relative to the WS Higgs field VEV. However we note that there is a finesse of our fit to the quark-lepton spectrum, according to which the right-handed component of the experimentally observed t-quark is actually the one having second generation $`SU(3)`$ quantum numbers and is thus really the proto-right-handed charm quark $`c_R`$. In a similar way the right-handed component of the experimentally observed charm quark has the third generation $`SU(3)`$ representation and is really the proto-right-handed top quark $`t_R`$. It is only the right-handed top and charm quarks that are permuted in this way, while for example the left-handed components are not. We have to make this identification of the proto-generation fields $`c_R`$ and $`t_R`$; otherwise we cannot suppress the $`b`$ quark and $`\tau `$ lepton masses. This is because, for the proto-fields, the charge differences between $`t_L`$ and $`t_R`$ are the same as between $`b_L`$ and $`b_R`$ and also between $`\tau _L`$ and $`\tau _R`$. So now it is simple to calculate the quantum numbers of the WS Higgs field $`\varphi _{WS}`$: $`\stackrel{}{Q}_{\varphi _{WS}}=\stackrel{}{Q}_{c_R}\stackrel{}{Q}_{t_L}`$ $`=`$ $`(0,{\displaystyle \frac{2}{3}},0,1)(0,0,{\displaystyle \frac{1}{6}},0)`$ (34) $`=`$ $`(0,{\displaystyle \frac{2}{3}},{\displaystyle \frac{1}{6}},1)`$ This means that the WS Higgs field will in fact be coloured under both $`SU(3)_2`$ and $`SU(3)_3`$. After breaking the symmetry down to the SM group, we will be left with the usual WS Higgs field of the SM and another scalar which will be an octet of $`SU(3)`$ and a doublet of $`SU(2)`$. This should not present any phenomenological problems, provided this scalar doesn’t cause symmetry breaking and doesn’t have a mass less than about 1 TeV. In particular an octet of $`SU(3)`$ cannot lead to baryon decay. In our model we take it that what in the Standard Model are seen as many very small Yukawa-couplings to the Standard Model Higgs field really represent chain Feynman diagrams, composed of propagators with Planck scale heavy particles (fermions) interspaced with order of unity Yukawa couplings to Higgs fields with the names $`W`$, $`T`$, $`\xi `$, and $`S`$, which are postulated to break the AGUT to the Standard Model Group. The small effective Yukawa couplings in the Standard Model are then generated as products of small factors, given by the ratios of the vacuum expectation values of $`W`$, $`T`$, and $`\xi `$ to the masses occurring in the propagators for the Planck scale fermions in the chain diagrams . The quantum numbers of our invented Higgs fields $`W`$, $`T`$, $`\xi `$ and $`S`$ are chosen—and it is remarkable that we succeeded so well—so as to make the order of magnitude for the suppressions of the mass matrix elements of the various mass matrices fit to the phenomenological requirements. After the choice of the quantum numbers for the replacement of the Weinberg Salam Higgs field in our model, eq. (34), the further quantum numbers needed to be picked out of the vacuum in order to give, say, mass to the b-quark is denoted by $`\stackrel{}{b}`$ and analogously for the other particles. For example: $`\stackrel{}{b}`$ $`=`$ $`\stackrel{}{Q}_{b_L}\stackrel{}{Q}_{b_R}\stackrel{}{Q}_{WS}`$ (35) $`\stackrel{}{c}`$ $`=`$ $`\stackrel{}{Q}_{c_L}\stackrel{}{Q}_{t_R}+\stackrel{}{Q}_{WS}`$ (36) $`\stackrel{}{\mu }`$ $`=`$ $`\stackrel{}{Q}_{\mu _L}\stackrel{}{Q}_{\mu _R}\stackrel{}{Q}_{WS}`$ (37) Here we denoted the quantum numbers of the quarks and leptons as e.g. $`\stackrel{}{Q}_{c_L}`$ for the left handed components of the proto-charmed quark. Note, as we remarked above, that $`\stackrel{}{c}`$ has been defined using the $`t_R`$ proto-field, since we have essentially swapped the right-handed charm and top quarks. Also the charges of the WS Higgs field are added rather than subtracted for up-type quarks. Next we attempted to find some Higgs field quantum numbers which, if postulated to have “small” vevs compared to the Planck scale masses of the intermediate particles, would give a reasonable fit to the order of magnitudes of the mass matrix elements. We were thereby led to the proposal: $$\stackrel{}{Q}_W=\frac{1}{3}(2\stackrel{}{b}+\stackrel{}{\mu })=(0,\frac{1}{2},\frac{1}{2},\frac{4}{3})$$ (38) $$\stackrel{}{Q}_T=\stackrel{}{b}\stackrel{}{Q}_W=(0,\frac{1}{6},\frac{1}{6},\frac{2}{3})$$ (39) $`\stackrel{}{Q}_\xi =\stackrel{}{Q}_{d_L}\stackrel{}{Q}_{s_L}`$ $`=`$ $`({\displaystyle \frac{1}{6}},0,0,0)(0,{\displaystyle \frac{1}{6}},0,0)`$ (40) $`=`$ $`({\displaystyle \frac{1}{6}},{\displaystyle \frac{1}{6}},0,0)`$ From the Fritzsch relation $`V_{us}\sqrt{\frac{m_d}{m_s}}`$ discussed in section 2, it is suggested that the two off-diagonal mass matrix elements connecting the d-quark and the s-quark be equally big. We achieve this approximately in our model by introducing a special Higgs field $`S`$, with quantum numbers equal to the difference between the quantum number differences for these 2 matrix elements in the down quark matrix. Then we postulate that this Higgs field has a vev of order unity in fundamental units, so that it does not cause any suppression but rather ensures that the two matrix elements get equally suppressed. Henceforth we will consider the vevs of the new Higgs fields as measured in Planck scale units and so we have: $$<S>=1$$ (41) and $`\stackrel{}{Q}_S`$ $`=`$ $`[\stackrel{}{Q}_{s_L}\stackrel{}{Q}_{d_R}][\stackrel{}{Q}_{d_L}\stackrel{}{Q}_{s_R}]`$ (42) $`=`$ $`({\displaystyle \frac{1}{6}},{\displaystyle \frac{1}{6}},0,1)`$ The existence of a non-suppressing field $`S`$ means that we cannot control phenomenologically when this $`S`$-field is used. Thus the quantum numbers of the other Higgs fields $`W`$, $`T`$, $`\xi `$ and $`\varphi _{WS}`$ given above have only been determined modulo those of the field $`S`$. ## 7 Quark and lepton mass matrices in AGUT We define the mass matrices by considering the mass terms in the SM to be given by: $$=\overline{Q}_LM_UU_R+\overline{Q}_LM_DD_R+\overline{L}_LM_EE_R+\mathrm{h}.\mathrm{c}.$$ (43) The mass matrices can be expressed in terms of the effective SM Yukawa matrices and the WS Higgs VEV by: $$M_f=Y_f\frac{<\varphi _{WS}>}{\sqrt{2}}$$ (44) We can now calculate the suppression factors for all elements in the Yukawa matrices, by expressing the charge differences between the left-handed and right-handed fermions in terms of the charges of the Higgs fields. They are given by products of the small numbers denoting the vevs of the fields $`W`$, $`T`$, $`\xi `$ in fundamental units and the order unity vev of $`S`$. In the following matrices we simply write $`W`$ instead of $`<W>`$ etc. for the vevs in Planck units. With the quantum number choice given above, the resulting matrix elements are—but remember that “random” complex order unity factors are supposed to multiply all the matrix elements—for the uct-quarks: $$Y_U\left(\begin{array}{ccc}SWT^2\xi ^2& WT^2\xi & W^2T\xi \\ SWT^2\xi ^3& WT^2& W^2T\\ S\xi ^3& 1& WT\end{array}\right)$$ (45) the dsb-quarks: $$Y_D\left(\begin{array}{ccc}SWT^2\xi ^2& WT^2\xi & T^3\xi \\ SWT^2\xi & WT^2& T^3\\ SW^2T^4\xi & W^2T^4& WT\end{array}\right)$$ (46) and the charged leptons: $$Y_E\left(\begin{array}{ccc}SWT^2\xi ^2& WT^2\xi ^3& S^2WT^4\xi \\ SWT^2\xi ^5& WT^2& S^2WT^4\xi ^2\\ S^3WT^5\xi ^3& W^2T^4& WT\end{array}\right)$$ (47) We can now set $`S=1`$ and fit the nine quark and lepton masses and three mixing angles, using 3 parameters: $`W`$, $`T`$ and $`\xi `$. That really means we have effectively omitted the Higgs field $`S`$ and replaced the maximal AGUT gauge group $`SMG^3\times U(1)_f`$ by the reduced AGUT group $`SMG_{12}\times SMG_3\times U(1)`$, which survives the spontaneous breakdown due to $`S`$. In order to find the best possible fit we must use some function which measures how good a fit is. Since we are expecting an order of magnitude fit, this function should depend only on the ratios of the fitted masses to the experimentally determined masses. The obvious choice for such a function is: $$\chi ^2=\left[\mathrm{ln}\left(\frac{m}{m_{\text{exp}}}\right)\right]^2$$ (48) where $`m`$ are the fitted masses and mixing angles and $`m_{\text{exp}}`$ are the corresponding experimental values. The Yukawa matrices are calculated at the fundamental scale which we take to be the Planck scale. We use the first order renormalisation group equations (RGEs) for the SM to calculate the matrices at lower scales. We cannot simply use the 3 matrices given by eqs. (45)–(47) to calculate the masses and mixing angles, since only the order of magnitude of the elements is defined. Therefore we calculate statistically, by giving each element a random complex phase and then finding the masses and mixing angles. We repeat this several times and calculate the geometrical mean for each mass and mixing angle. In fact we also vary the magnitude of each element randomly, by multiplying by a factor chosen to be the exponential of a number picked from a Gaussian distribution with mean value 0 and standard deviation 1. We then vary the 3 free parameters to find the best fit given by the $`\chi ^2`$ function. We get the lowest value of $`\chi ^2`$ for the VEVs: $`W`$ $`=`$ $`0.179`$ (49) $`T`$ $`=`$ $`0.071`$ (50) $`\xi `$ $`=`$ $`0.099`$ (51) The result of the fit is shown in table 1. This fit has a value of: $$\chi ^2=1.87$$ (52) This is equivalent to fitting 9 degrees of freedom (9 masses + 3 mixing angles - 3 Higgs vevs) to within a factor of $`\mathrm{exp}(\sqrt{1.87/9})1.58`$ of the experimental value. This is better than might have been expected from an order of magnitude fit. We can also fit to different experimental values of the 3 light quark masses by using recent results from lattice QCD, which seem to be consistently lower than the conventional phenomenological values. The best fit in this case is shown in table 2. The corresponding values of the Higgs vevs are: $`W`$ $`=`$ $`0.123`$ (53) $`T`$ $`=`$ $`0.079`$ (54) $`\xi `$ $`=`$ $`0.077`$ (55) and this fit has a larger value of: $$\chi ^2=3.81$$ (56) But even this is good for an order of magnitude fit. ## 8 Neutrino mass and mixing Physics beyond the SM can generate an effective light neutrino mass term $$_{\nu mass}=\underset{i,j}{}\psi _{i\alpha }\psi _{j\beta }ϵ^{\alpha \beta }(M_\nu )_{ij}$$ (57) in the Lagrangian, where $`\psi _{i,j}`$ are the Weyl spinors of flavour $`i`$ and $`j`$, and $`\alpha ,\beta =1,2`$. Fermi-Dirac statistics means that the mass matrix $`M_\nu `$ must be symmetric. In models with chiral flavour symmetry we typically expect the elements of the mass matrices to have different orders of magnitude. The charged lepton matrix is then expected to give only a small contribution to the lepton mixing. As a result of the symmetry of the neutrino mass matrix and the hierarchy of the mass matrix elements it is natural to have an almost degenerate pair of neutrinos, with nearly maximal mixing . This occurs when an off-diagonal element dominates the mass matrix. A neutrino mass matrix of this texture is generated in the AGUT model, by tree level diagrams involving the exchange of two Weinberg Salam Higgs tadpoles and the appropriate combination of Planck scale Higgs field tadpoles. The combination which leads to the mass term $`(M_\nu )_{ij}`$ between $`\nu _{Li}`$ and $`\nu _{Lj}`$ is determined by the equation $$\left(\stackrel{}{Q}_\theta \right)_{ij}=\stackrel{}{Q}_{\nu Li}+\stackrel{}{Q}_{\nu Lj}+2\stackrel{}{Q}_{\varphi _{WS}}$$ (58) Here the sum is over the charge vectors for the combination of Planck scale Higgs fields ($`W`$, $`T`$, $`\xi `$ and $`S`$) exchanged. In this way we obtain the neutrino mass matrix $$M_\nu \frac{\varphi _{WS}^2}{M_{Pl}}\left(\begin{array}{ccc}W^2\xi ^4T^4& W^2\xi T^4& W^2\xi ^3T\\ W^2\xi T^4& WT^5& W^2T\\ W^2\xi ^3T& W^2T& W^2T^2\xi ^2\end{array}\right)$$ (59) where we have set $`<S>=1`$. The off-diagonal element $`(M_\nu )_{23}=(M_\nu )_{32}`$ clearly dominates this matrix, so that we have large mu-tau mixing (between the nearly degenerate mass eigenstates $`\nu _2`$ and $`\nu _3`$). The mixing matrix $`U_\nu `$ is given by $$U_\nu \left(\begin{array}{ccc}1& \frac{\xi ^3}{\sqrt{2}}& \frac{\xi ^3}{\sqrt{2}}\\ \xi ^3& \frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}\\ \xi T^3& \frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}\end{array}\right)$$ (60) We also have the ratio of neutrino mass squared differences $$\frac{\mathrm{\Delta }m_{23}^2}{\mathrm{\Delta }m_{12}^2}2T\xi ^21.4\times 10^3$$ (61) giving a hierarchy that is not suitable for the simultaneous solution of the solar and atmospheric neutrino problems. In any case, the mass scale is much too small to give suitable masses for the atmospheric neutrino problem. This is because, even if the $`(M_\nu )_{23}`$ element was unsuppressed by Planck scale Higgs vevs, the see-saw mass $$\frac{<\varphi _{WS}>^2}{M_{Planck}}3\times 10^6\text{eV}$$ (62) would still be too small. So, it is necessary to introduce a new mass scale into the AGUT model in order to obtain observable neutrino masses and mixings. This may be done by extending the AGUT Higgs spectrum to include a weak isotriplet Higgs field $`\mathrm{\Delta }`$ with SM weak hypercharge $`\frac{y}{2}=1`$. However there is some unnaturalness in obtaining a value for $`<\mathrm{\Delta }^0>`$ from the scalar potential some orders of magnitude greater than the see-saw mass of eq. (62) Furthermore we need extra structure for the lepton mass matrices and must relax the assumption that all the independent matrix elements are of different orders of magnitude. For example $`M_\nu `$ may have two order of magnitude degenerate elements $`AB`$ with a texture of the form: $$M_\nu =\left(\begin{array}{ccc}\times & A& B\\ A& \times & \times \\ B& \times & \times \end{array}\right)$$ (63) where $`\times `$ indicates texture zeros. The mass eigenvalues are given by: $$m_{\nu i}=\pm \sqrt{A^2+B^2},\mathrm{\hspace{0.17em}0},(i=1,2,3)$$ (64) although these will be slightly altered when the effects of the small elements represented by texture zeros are included. With these eigenvalues we clearly have a hierarchy in $`\mathrm{\Delta }m^2`$’s with the more degenerate pair being heavier: $$\mathrm{\Delta }m_{12}^2\mathrm{\Delta }m_{13}^2\mathrm{\Delta }m_{23}^2.$$ (65) So we take $`\mathrm{\Delta }m_{12}^2=\mathrm{\Delta }m_{solar}^2`$, $`\mathrm{\Delta }m_{23}^2=A^2+B^210^3\text{eV}^2`$, where $`\mathrm{\Delta }m_{solar}^2`$ will depend on the type of solution we adopt for the solar neutrinos. The corresponding neutrino mixing matrix (assuming that the charged lepton mass matrix $`M_E`$ is quasi-diagonal) is: $`U_\nu `$ $``$ $`\left(\begin{array}{ccc}1& 0& 0\\ 0& \mathrm{cos}\theta & \mathrm{sin}\theta \\ 0& \mathrm{sin}\theta & \mathrm{cos}\theta \end{array}\right)\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}& 0\\ \frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}& 0\\ 0& 0& 1\end{array}\right)`$ (72) $`=`$ $`\left(\begin{array}{ccc}\frac{1}{\sqrt{2}}& \frac{1}{\sqrt{2}}& 0\\ \frac{1}{\sqrt{2}}\mathrm{cos}\theta & \frac{1}{\sqrt{2}}\mathrm{cos}\theta & \mathrm{sin}\theta \\ \frac{1}{\sqrt{2}}\mathrm{sin}\theta & \frac{1}{\sqrt{2}}\mathrm{sin}\theta & \mathrm{cos}\theta \end{array}\right)`$ (76) where $$\mathrm{tan}\theta =\frac{B}{A}.$$ (77) From the first row we can see that $`\nu _e`$ is maximally mixed between $`\nu _1`$ and $`\nu _2`$, so that its mixing does not contribute to the atmospheric neutrino anomaly, and there will be no effect observable at Chooz. The atmospheric neutrino anomaly will be entirely due to large $`\nu _\mu \nu _\tau `$ mixing and, in order that the mixing be large enough, we need $`\mathrm{sin}^22\theta 0.8`$ ($`90\%C.L`$) which requires $$0.56\frac{B}{A}1.8$$ (78) so that, although $`A`$ and $`B`$ must be order of magnitude degenerate, it is not necessary to do any fine tuning. The solar neutrino problem is explained by vacuum oscillations, although whether it is an ‘energy-independent’ or a ‘just-so’ solution will depend on the small elements which we have neglected. It cannot be explained by an MSW type solution since the mixing between $`\nu _e`$ and $`\nu _\mu `$ is too large for this type of solution, and will remain too large even after the texture zeroes are removed. The particular case of $`B=A`$ for this texture corresponds to the popular bi-maximal mixing solution to the solar and atmospheric neutrino problems. This type of structure cannot explain the LSND result and does not give a significant contribution to hot dark matter, since the sum of the neutrino masses is given by $`{\displaystyle m_\nu }`$ $``$ $`2\sqrt{A^2+B^2}2\sqrt{\mathrm{\Delta }m_{atm}^2}`$ (79) $`<`$ $`0.2\text{eV}`$ We have not been able to extend the Higgs sector of the AGUT model in such a way as to obtain a neutrino mass matrix $`M_\nu `$ with the above texture of eq. (63). However we have constructed an anomaly free Abelian extension of the Standard Model, which naturally yields a mass matrix $`M_\nu `$ of this type. This $`SMG\times U(1)^2`$ model was inspired by the AGUT model and has exactly the same charged fermion spectrum as in the AGUT fit of Table 1. In order to rescue the AGUT neutrino mass and mixing predictions, it seems necessary to introduce yet another Higgs field and obtain the large mixing required for the atmospheric neutrino problem from the charged lepton mass matrix $`M_E`$. The solution to the solar neutrino problem can then be obtained from $`M_\nu `$ or from the mixing due to small elements in $`M_E`$. This, of course, has to be achieved without signicantly disturbing the quality of the AGUT fit to the charged fermion spectrum. ## 9 Conclusions We emphasized the hierarchical structure of the quark-lepton mass spectrum and how it points to a mass protection mechanism, controlled by approximately conserved chiral (gauge) charges beyond the Standard Model. The structure of ansätze for the fermion mass matrices, suggested by the hierarchy of masses and mixing angles, was briefly discussed. A recent ansatz based on a lightest flavour mixing mechanism was discussed, which gives simple and compact formulae for all the CKM mixing angles in terms of the quark masses. The anti-grand unification theory (AGUT), and how the associated multiple point principle (MPP) is used to predict the values of the three Standard Model gauge coupling constants, was described. Applied to the case of the pure Standard Model, the MPP leads to our predictions for the top quark and Higgs pole masses: $`M_t=173\pm 5`$ GeV and $`M_H=135\pm 9`$ GeV. The AGUT group $`SMG^3\times U(1)_f`$ is characterised by being the largest anomaly-free gauge group acting on just the 45 SM Weyl fermions, without any unification of the SM irreducible representations. This group assigns a unique set of anomaly free chiral gauge charges to the quarks and leptons. With an appropriate choice of Higgs field quantum numbers, the AGUT chiral charges naturally give a realistic charged fermion mass hierarchy. An order of magnitude fit in terms of 3 Higgs vevs is given in Table 1, which reproduces all the masses and mixing angles within a factor of two. The most characteristic feature of the fit is that, apart from the $`t`$ and $`c`$ quarks, the masses of the particles in the same generation are predicted to be degenerate (but only in order of magnitude) at the Planck scale. The worst feature is the deviation, by a factor of about 2, between the fitted and experimental values for $`m_s`$ and $`V_{cb}`$. On the other hand, the puzzle of the neutrino masses and mixing angles presents a challenge to the model. It is necessary to introduce a new mass scale into the AGUT model, using say a weak isotriplet Higgs field $`\mathrm{\Delta }`$, in order to generate a neutrino mass appropriate to atmospheric neutrino oscillations. Using a reduced model, based on the gauge group $`SMG\times U(1)^2`$, it is possible to obtain a reasonably natural solution to the solar and atmospheric neutrino problems and, at the same time, reproduce the successful AGUT fit to the charged fermion spectrum. However it is not possible to embed this Abelian extension of the SM into the AGUT, since one cannot choose a consistent set of non-Abelian representations for the Higgs fields. It appears that we shall have to relax the assumption that the charged lepton mass matrix is quasi-diagonal, in order to rescue the AGUT model. ## 10 Acknowledgements I should like to thank the organisers George Koutsoumbas, Nick Tracas and George Zoupanos for their hospitality in Corfu.
no-problem/9908/hep-th9908020.html
ar5iv
text
# Untitled Document HUTP-99/A037, DAMTP-1999-98 hep-th/9908020 Instantons and Non-renormalisation in AdS/CFT Rajesh Gopakumar<sup>1</sup> gopakumr@tomonaga.harvard.edu Lyman Laboratory of Physics, Harvard University Cambridge, MA 02138, USA Michael B. Green<sup>2</sup> M.B.Green@damtp.cam.ac.uk DAMTP, Silver Street, Cambridge CB3 9EW, UK The series of perturbative fluctuations around a multi-instanton contribution to a specific class of correlation functions of supercurrents in $`𝒩=4`$ supersymmetric $`SU(N)`$ Yang–Mills theory is examined in the light of the AdS/CFT correspondence. Subject to certain plausible assumptions, we argue that a given term in the $`1/N`$ expansion in such a background receives only a finite number of perturbative corrections in the ’t Hooft limit. Such instanton non-renormalisation theorems would explain, for example, the exact agreement of certain weak coupling Yang–Mills instanton calculations with the strong coupling predictions arising from D-instanton effects in string theory amplitudes. These non-renormalisation theorems essentially follow from the assumption of a well defined derivative $`(\alpha ^{})`$ expansion in the string theory dual of the Yang–Mills theory. August 1999 1. Introduction The conjectured equivalence of type IIB superstring theory on $`AdS_5\times S^5`$ to the boundary $`𝒩=4`$ supersymmetric $`SU(N)`$ Yang–Mills conformal field theory \[1,,2,,3\] has been tested by a variety of calculations at leading order in the large-$`N`$ limit and at large values of the ’t Hooft coupling, $`\lambda =g_{_{YM}}^2N/4\pi `$ ($`g_{_{YM}}^2`$ is the Yang–Mills coupling constant). Many of these tests, such as those of certain two and three point correlation functions, have relied on non-renormalisation theorems/conjectures \[4\] and therefore allow the meaningful comparison of the regimes of strong and weak ’t Hooft coupling. Clearly, the ideal way of developing the AdS/CFT correspondence beyond the limited large $`\lambda `$ region in which it has so far been studied would be to explicitly quantize IIB superstring theory in an $`AdS_5\times S^5`$ background. Unfortunately, this is a daunting problem, even at tree level – in part because of the presence of a nonzero condensate of $`RR`$ background fields associated with the nonzero $`F_5`$ flux. In the absence of an explicit construction of string amplitudes most concrete calculations have made use of known low order terms in the expansion of the effective supergravity action in powers of the dimensionless parameter $`\alpha ^{}/L^2`$ ($`L`$ is the size of the $`AdS_5`$ and $`S^5`$ background and $`\alpha _{}^{}{}_{}{}^{1/2}`$ is the string distance scale). In fact, knowing the complete effective action for the massless fields of string theory would be sufficient to compute the Yang-Mills correlation functions of the relevant dual operators, but we are far from achieving this. Nevertheless, in the following we will show how some reasonable assumptions concerning the structure of the low energy expansion of type IIB string theory lead to a number of non-renormalisation theorems in the instanton sector of certain Yang-Mills correlation functions. To be concrete, we will consider the D-instanton contributions to the four graviton scattering amplitude and arrive at statements regarding the corresponding Yang-Mills instanton terms in the AdS/CFT dual correlation functions of four energy-momentum tensors. Similar statements also apply to any of the Yang-Mills correlations functions that are related by supersymmetry. Specifically, what we will see is that, for certain ‘protected’ parts of the correlation functions, the ’t Hooft expansion around an instanton background has only a finite number of perturbative terms in $`\lambda `$ at each order in $`1/N`$. In particular, we will see that at leading order in $`N`$ only the $`\lambda `$-independent semi-classical term arises. This would account for the precise agreement (for any instanton number $`K`$, at leading order in $`N`$) between the D-instanton contributions, at leading order in the $`\alpha ^{}`$ (or $`1/\lambda `$) expansion \[5\], with the semi-classical (small-$`\lambda `$) contributions of Yang–Mills instantons \[6,,7\]. In the case of two and three point functions the space-time dependence is completely determined by (super) conformal invariance. However, the matching of the string theory D-instanton and Yang–Mills instanton contributions to the protected correlation functions involves matching non-trivial functions of the space-time positions (functions of two independent cross ratios in the case of the correlator of four stress tensors), together with specific dependence on $`N`$, $`\lambda `$ and the instanton number. 1.1. Overview of the Correspondence The AdS/CFT conjecture \[1\] gives a relation between the parameters of the string theory – the dimensionless $`AdS_5\times S^5`$ scale $`L^2/\alpha ^{}`$, the $`RR`$ scalar field, $`C^{(0)}`$, and the coupling constant $`g=e^\varphi =\tau _2^1`$ – and those of the Yang–Mills theory with gauge group $`SU(N)`$. $$g=\frac{g_{_{YM}}^2}{4\pi },2\pi C^{(0)}=\theta ,\frac{L^4}{\alpha _{}^{}{}_{}{}^{2}}=g_{_{YM}}^2N4\pi \lambda ,$$ where $`\theta `$ is the constant axionic angle. This means that the constant value of the complex coupling constant, $`\tau \tau _1+i\tau _2=C^{(0)}+ie^\varphi `$, in the $`AdS_5\times S^5`$ background is identified with the complex Yang–Mills coupling, $$\tau =\frac{\theta }{2\pi }+i\frac{4\pi }{g_{_{YM}}^2}.$$ In the following, $`\tau `$ will always be assumed to be equal to this constant value (mostly with $`\tau _2^1<<1`$). According to the prescription of \[2,,3\] the amplitudes of the bulk superstring theory in the $`AdS_5\times S^5`$ background with fields propagating to specified values at points on the boundary are equivalent to correlation functions of composite operators in the boundary Yang–Mills theory. The boundary values of the bulk fields are interpreted as sources coupling to the operators in the Yang–Mills theory. Collectively denoting the independent cross ratios of the positions of the boundary fields by $`\eta `$, the resulting amplitude for an $`n`$-point function in the gauge theory can be written as a finite sum over contributions of the form $$_n^s(\frac{\alpha ^{}}{L^2},\tau ,\overline{\tau },\eta )A_n=_n^{YM}(\lambda ,N,\theta ,\eta )A_n.$$ The right-hand side is just a rewriting in terms of the variables $`N`$, $`\lambda `$ and $`\theta `$ in which it is natural to express the correlation functions of the Yang–Mills theory. The finite sum involves factors $`A_n`$ which span an independent set of tensor structures consistent with the space-time quantum numbers and symmetries of the $`n`$-point functions. Moreover, the $`A_n`$’s all have a common factor, also dictated by symmetry (a function of the space-time separations $`|x_ix_j|`$ ), which carries the dimension of the correlation function. Therefore, almost all the non-trivial information about the correlation functions really lie in the functions $`_n`$. In the following, we will be exclusively concerned with the large $`N`$ limit of (1.1). The expansion of gauge theory amplitudes in $`1/N`$ translates into a small $`g`$ expansion in the left-hand side of (1.1). The ’t Hooft expansion of Yang–Mills amplitudes takes the familiar form (for convenience, we will drop the subscript $`n`$ in much that follows), $$^{^{YM}}(\lambda ,N,\theta ,\eta )=N^2\left[_0^{^{YM}}(\lambda ,\eta )+\frac{1}{N^2}_1^{^{YM}}(\lambda ,\eta )+\mathrm{}+\frac{1}{N^{2k}}_k^{^{YM}}(\lambda ,\eta )+\mathrm{}\right]+\mathrm{}.$$ Here the second ellipsis includes Yang–Mills instanton terms of the form $`e^{2\pi |K|\frac{N}{\lambda }+iK\theta }`$, each coming with its series of fluctuations. Although these instanton terms are exponentially suppressed they are uniquely specified by their phase. It is the structure of the ’t Hooft expansion of fluctuations around a particular instanton background that we will focus on in Sec. 2. In the correspondence with the dual string theory, the terms in (1.1) which are powers of $`1/N`$ arise from perturbative string contributions with $`k`$ being the world sheet genus. The instanton terms which are suppressed by powers of $`e^N`$ are non-perturbative and can be identified with D-instanton contributions \[8\]. Though the form of the $`1/N`$ expansion in (1.1) was originally motivated by weakly coupled perturbation theory, the existence of the dual string theory with the identifications (1.1) implies such a form should apply for all $`\lambda `$. In particular, we will exploit the existence of a well defined expansion for large $`\lambda `$ that is defined by the $`\alpha ^{}`$ expansion of the string theory. 2. Instanton Non-renormalisation Theorems The instanton calculations in \[6,,7\] involved correlation functions of various combinations of the superconformal currents that make up a short (256 component) $`𝒩=4`$ supermultiplet. However, the general structure of interest to us does not depend on which of these correlation functions is considered so we will focus on a specific tensor structure in the correlation function of four stress tensors. This particular tensor structure can be defined by its relation, via the AdS/CFT correspondence, to the $`^4`$ term in the type IIB effective action (where $``$ denotes the Weyl curvature). This ten-dimensional term has the tensor structure $$^4t^{M_1\mathrm{}M_8}t_{N_1\mathrm{}N_8}R_{M_1M_2}^{N_1N_2}\mathrm{}R_{M_7M_8}^{N_7N_8}$$ with $`t`$ being a standard eighth-rank tensor. We wish to consider the linearization of the four curvatures around the $`AdS_5\times S^5`$ background, keeping only the polarisations in the $`AdS_5`$ directions. The four-graviton scattering amplitude is expressed as a functional of the boundary ($`S^4`$) values of the graviton by attaching a spin-two bulk-to-boundary propagator to each graviton leg in the linearized vertex. Since the boundary graviton field is interpreted in the Yang–Mills theory as the source for the stress tensor, this procedure defines a particular tensor contribution to the correlation function of four stress tensors \[5\] which can be expressed as $$\frac{L^2}{\alpha ^{}}\tau _2^{1/2}f_1^{(0,0)}(\tau ,\overline{\tau })g_1(\eta )A_4.$$ The $`\alpha _{}^{}{}_{}{}^{1}`$ dependence reflects the fact that the $`^4`$ term is of order $`\alpha _{}^{}{}_{}{}^{3}`$ relative to the Einstein–Hilbert term. The function $`A_4`$ has, in addition to the particular tensor structure determined by the bulk-boundary correspondence described above, a factor of $`_{i<j}|x_ix_j|^{\frac{8}{3}}`$, which is fixed by conformal invariance. The residual dependence on the positions of the boundary operators is contained in the function of the two independent cross-ratios which has been denoted as $`g_1(\eta )`$. Explicit expressions for $`g_1(\eta )`$ in the case of closely related four-point functions were obtained in \[6,,9\]. The modular invariant function of the (complex) string coupling, $`f_1^{(0,0)}(\tau ,\overline{\tau })`$, is a nonholomorphic Eisenstein series that has the Fourier expansion in powers of $`e^{2\pi i\tau _1}`$ \[8\], $$\begin{array}{cc}\hfill f_1^{(0,0)}(\tau ,\overline{\tau })& \underset{(m,n)(0,0)}{}\frac{\tau _2^{3/2}}{|m+n\tau |^3}=\underset{K=\mathrm{}}{\overset{\mathrm{}}{}}_K^1(\tau _2)e^{2\pi iK\tau _1}\hfill \\ & =2\zeta (3)\tau _2^{\frac{3}{2}}+\frac{2\pi ^2}{3}\tau _2^{\frac{1}{2}}+4\pi \underset{K=1}{\overset{\mathrm{}}{}}|K|^{1/2}\mu (K,1)\hfill \\ & \times \left(e^{2\pi iK\tau }+e^{2\pi iK\overline{\tau }}\right)\left(1+\underset{k=1}{\overset{\mathrm{}}{}}(4\pi K\tau _2)^k\frac{\mathrm{\Gamma }(k1/2)}{\mathrm{\Gamma }(k1/2)k!}\right).\hfill \end{array}$$ Here $`\mu (K,1)=_{d|K}d^2`$. The $`K=0`$ term contains the perturbative tree-level and one-loop contributions while the $`K0`$ terms are D-instanton contributions. The leading $`\tau _2`$ independent term in the charge-$`K`$ D-instanton sector was found to agree with a weak coupling Yang-Mills calculation in \[7\] (at least for the related sixteen-dilatino correlation function). More generally, in the $`AdS_5\times S^5`$ background there will be contributions proportional to $`A_4`$ which are of higher order in $`\alpha ^{}`$. These come from higher derivative terms in the effective action and can be studied in a Taylor expansion for small $`\alpha ^{}/L^2`$, $$^s(\frac{\alpha ^{}}{L^2},\tau ,\overline{\tau },\eta )A_4=\underset{l=1}{}\left(\frac{\alpha ^{}}{L^2}\right)^{l2}F_l(\eta ,\tau ,\overline{\tau })A_4,$$ where the $`^4`$ contribution is the first ($`l=1`$) term in the series (so that $`F_1(\eta ,\tau ,\overline{\tau })=g_1(\eta )\tau _2^{1/2}f_1^{(0,0)}(\tau ,\overline{\tau })`$). Examples of higher derivative terms in the ten dimensional effective action that would contribute to (2.1) include terms of the general form (in string frame) $$(\alpha ^{})^{2k3}d^{10}x\sqrt{G^{(10)}}e^{(5k11/2)\varphi }F_5^{4k4}^4f_k^{(0,0)}(\tau ,\overline{\tau }).$$ The modular functions $`f_k^{(0,0)}`$ that appear here, have been conjectured to be generalised Eisenstein functions \[10,,11\]. In the $`AdS_5\times S^5`$ background with its constant five-form field strength $`F_5`$, these terms can give contributions proportional to $`A_4`$. Another class of terms suggested in \[12,,13\] involve derivatives acting on $`^4`$, which may also give nonzero contributions $`A_4`$ in the $`AdS_5\times S^5`$ background. There might also be terms that contribute to $`A_4`$ that cannot be expressed in terms of a local ten-dimensional action. What about the possibility of terms which are non-perturbative in $`\alpha ^{}`$, which are exponentially suppressed in the Taylor expansion, such as $`e^{L^2/\alpha ^{}}`$? For small $`\alpha ^{}/L^2`$, we understand terms of this type as coming from non-trivial saddle points of the world sheet theory, namely world-sheet instantons. But there are no topologically non-trivial two-cycles for the world-sheet to wrap in $`AdS_5\times S^5`$ so such terms cannot appear in the string genus expansion.<sup>3</sup> In other instances of large $`N`$ gauge theories dual to closed strings, world sheet instantons are present and play an important role \[14\]. One might argue that this does not rule out $`e^{L^2/\alpha ^{}}`$ contributions associated with non-perturbative $`\tau `$ dependence. But this also seems unlikely since we understand such non-perturbative terms as coming from D-instantons and again there is no obvious origin for world-sheet instanton contributions in a D-instanton background. We will therefore make the ansatz that (2.1) is the complete expression for the coefficient of $`A_4`$ in the correlation function of four energy-momentum tensors, at least for sufficently small $`\alpha ^{}/L^2`$. In practice, we will only need the weaker assumption that this is so for the D-instanton contributions to (2.1) (the terms with phases $`e^{2\pi iK\tau _1}`$, $`|K|1`$). Our arguments will show that this assumption is, at least, self consistent. The $`SL(2,Z)`$ duality symmetry of the IIB theory is related via the AdS/CFT correspondence to the Montonen–Olive duality of $`𝒩=4`$ supersymmetric Yang–Mills theory. This requires that $`F_l(\tau ,\overline{\tau },\eta )`$ has specific modular properties which means that it has the form, $$F_l(\tau ,\overline{\tau },\eta )=\tau _2^{1\frac{l}{2}}H_l(\tau ,\overline{\tau },\eta )$$ where $`H_l(\tau ,\overline{\tau },\eta )`$ is modular invariant (a scalar under $`SL(2,Z)`$). The explicit power of $`\tau _2^{1\frac{l}{2}}`$ arises from the transformation in the effective action from the string to the Einstein frame (where the metric is $`SL(2,Z)`$ neutral) due to the factors of the metric appearing with the powers of $`\alpha ^{}/L^2`$. In terms of Yang-Mills variables, this is easy to see since the coefficient of $`H_l`$ in (2.1) is the combination (from (2.1) and (2.1)) $$\left(\frac{\alpha ^{}}{L^2}\right)^{l2}\tau _2^{1\frac{l}{2}}=\frac{1}{\lambda ^{\frac{l}{2}1}}g_s^{\frac{l}{2}1}=N^{\frac{l}{2}+1},$$ which is inert under $`SL(2,Z)`$. Now consider the non-perturbative part of the modular function $`H_l(\tau ,\overline{\tau },\eta )`$ coming from BPS charge-$`K`$ D-instantons. This amounts to picking out the saddle point (when $`g=\tau _2^1<<1`$) with the exponential $`e^{2\pi (iK\tau _1|K|\tau _2)}`$ dependence.<sup>4</sup> In the charge-$`K`$ sector with phase $`e^{2\pi iK\tau _1}`$ there could be contributions from non-BPS configurations of $`K+K^{}`$ instantons and $`K^{}`$ anti-instantons. These would be suppressed by an additional factor of $`e^{4\pi K^{}N/\lambda }`$. Due to their different $`N`$ dependence such terms would not enter our considerations, even if they were present. We expect this to take the generic form, $$H_l(\tau ,\overline{\tau },\eta )|_K=d(K,l)e^{2\pi (|K|\tau _2iK\tau _1)}[h_0^{(l)}(\eta )+\tau _2^1h_1^{(l)}(\eta )+\tau _2^2h_2^{(l)}(\eta )+\mathrm{}],$$ an expression which deserves further explanation. Firstly, the successive terms in this series are spaced in integer powers of $`\tau _{2}^{}{}_{}{}^{1}`$ since they arise in string theory from world-sheet configurations with increasing numbers of boundaries with Dirichlet conditions and/or handles. The functions $`h_i^{(l)}(\eta )`$ (which also depend on the instanton charge $`K`$), that appear here, are severely constrained by the fact that $`H_l`$ is a modular function. Secondly, we have assumed that there is no $`\tau _2=g^1`$ dependence in the overall factor $`d(K,l)`$. Any such dependence would have to be a power that arises from the zero mode integrations around the D-instanton. This power should not depend on $`l`$. But we know from the $`l=1`$ case (2.1) that there is no such overall factor.<sup>5</sup> Actually, the essence of our conclusion will not be affected even if we had an overall factor of $`g^{n_l}`$, with $`n_l`$ taking values over non-negative integers. This statement should also be a consequence of supersymmetric cancellations of bosonic zero mode contributions with fermionic ones. Therefore, this might be special to a class of correlation functions such as the ones in the short multiplet we are concerned with. Rewriting the amplitude in terms of $`N`$, $`\lambda `$ and $`\theta `$ and summing the total contribution to the $`K`$ instanton background from all powers in the derivative expansion (all $`l`$) gives (using (2.1),(2.1) and (2.1)) $$\begin{array}{cc}\hfill ^{YM}|_K=& e^{2\pi \frac{|K|N}{\lambda }+iK\theta }\hfill \\ & \underset{l=1}{}d(K,l)N^{1\frac{l}{2}}\left[h_0^{(l)}(\eta )+\left(\frac{\lambda }{N}\right)h_1^{(l)}(\eta )+\left(\frac{\lambda }{N}\right)^2h_2^{(l)}(\eta )+\mathrm{}\right].\hfill \end{array}$$ This can be reorganized into a ’t Hooft expansion by grouping together the powers of $`1/N`$, which gives the perturbation expansion around the BPS $`K`$-instanton configuration, $$^{YM}|_K=e^{2\pi \frac{|K|N}{\lambda }+iK\theta }\underset{m=1}{}N^{1\frac{m}{2}}f_m(\lambda ,\eta ).$$ At first sight it might seem surprising that the series of fluctuations about an instanton should have a ’t Hooft expansion in powers of $`N^{1/2}`$. From the string viewpoint the spacing by half-integer powers of $`1/N`$ arises very naturally from the presence of all integer powers of $`\alpha ^{}`$ in (2.1). Furthermore, this feature is confirmed directly in the Yang–Mills theory by a saddle point analysis of the contribution of the exact zero-mode measure (as for example in (5.7) of \[7\]). For general instanton number $`K>1`$ the fluctuations around the saddle point are in powers of $`N^{1/2}`$. The series of fractional powers of $`N`$ therefore arises from the $`K`$-instanton measure. Only in the case $`K=1`$ does the series consist of terms with integer spaced powers of $`1/N`$ (starting with $`N^{1/2}`$). The key point following from the structure of (2.1) is that the power of $`\lambda `$ in the expansion is always bounded by that of $`N`$. In other words, each function $`f_m(\lambda ,\eta )`$ in (2.1) is a polynomial in $`\lambda `$, $$f_m(\lambda ,\eta )=\underset{k=0}{\overset{[\frac{m1}{2}]}{}}\lambda ^kh_k^{(m2k)}(\eta ).$$ For example, $$f_1(\lambda ,\eta )=h_0^{(1)}(\eta ),f_2(\lambda ,\eta )=h_0^{(2)}(\eta ),f_3(\lambda ,\eta )=h_0^{(3)}(\eta )+\lambda h_1^{(1)}(\eta ),\mathrm{}.$$ By our initial arguments, this is the complete form of the answer for small $`\alpha ^{}`$ or equivalently large $`\lambda `$. Barring the (unlikely) possibility of a phase transition as a function of $`\lambda `$, the knowledge that $`f_m`$ is a polynomial in positive powers of $`\lambda `$ at large $`\lambda `$ allows it to be analytically continued to weak ’t Hooft coupling.<sup>6</sup> Even though the orginal expansion in $`\alpha ^{}`$ (2.1) and $`g`$ (2.1) may only be asymptotic, we can nevertheless trust (2.1) to small $`\lambda `$. This is similar to the statement that though the full perturbative expansion in gauge theories has zero radius of convergence, the planar diagram expansion can be trusted in some finite radius . But then the polynomial form of $`f_m`$ means that there are only a finite number of terms in the small-$`\lambda `$ perturbation expansion for each power of $`N`$. Thus the very structure of the string expansion implies a sequence of non-trivial non-renormalisation theorems for ’t Hooft perturbation theory around an instanton background. In particular, the leading large $`N`$ term comes from $`m=1`$ and that is just a constant as far as its dependence on $`\lambda `$ is concerned. In other words, it receives only a semi-classical contribution. This ‘explains’ the fact that the semi-classical approximation to the $`K`$-instanton contribution to $`𝒩=4`$ $`SU(N)`$ Yang–Mills theory at leading order in $`N`$ (the term of order $`N^{1/2}`$ which was evaluated in \[7\]) agrees precisely with the expression predicted by the AdS/CFT correspondence \[5\]. Moreover, from (2.1) we see that the next to leading term should behave as $`N^0h_0^{(2)}(\eta )`$ and is also independent of $`\lambda `$ and therefore semi-classically exact. Verifying this prediction would require knowledge of the $`l=2`$ term in (2.1) that contributes at order $`(\alpha ^{})^0`$. Other predictions can also be tested. For instance, the structure of $`f_3`$ in (2.1) requires that the next term in the $`1/N`$ expansion has only two terms in the loop expansion. Also, in the special case of instanton number $`K=1`$ the semi-classical contribution was computed in for all values of $`N`$. In this case the expansion is in integer powers of $`1/N`$, which immediately determines that $$h_0^{(2k)}(\eta ,K=1)=0,h_0^{(2k1)}(\eta ,K=1)=b_kg_1(\eta )$$ where $`b_k`$ are the coefficients in the exact answer \[16\], $$\frac{\mathrm{\Gamma }(N\frac{1}{2})}{\mathrm{\Gamma }(N1)}=N^{\frac{1}{2}}\underset{l=1}{\overset{\mathrm{}}{}}b_kN^{k+1}=N^{\frac{1}{2}}\left(1\frac{5}{8}\frac{1}{N}\frac{23}{128}\frac{1}{N^2}+\mathrm{}\right).$$ 3. Comments and Conclusions We have seen by making rather minimal assumptions that certain instanton contributions in the ’t Hooft limit of $`𝒩=4`$ supersymmetric $`SU(N)`$ Yang–Mills theory receive only a finite number of perturbative corrections at a given order in the $`1/N`$ expansion. More precisely, at order $`N^{1\frac{m}{2}}`$ there are $`[\frac{m+1}{2}]`$ terms in the power series in the ’t Hooft coupling, $`\lambda `$, starting with $`\lambda ^0`$. It should be emphasised that, unlike with usual non-renormalisation theorems, our statements only apply at each order in $`1/N`$ in the ’t Hooft limit whereas for finite $`N`$ perturbative terms appear at all loops. Although our arguments do not make direct use of supersymmetry, this enters indirectly since the AdS/CFT correspondence does require supersymmetry. This is similar in spirit to the way in which the mere existence of a Lorentz invariant eleven-dimensional limit of M-theory implies non-trivial facts about D0-brane quantum mechanics, as in the DLCQ description. The assumptions we have made have the virtue that they can be checked by direct evaluation of perturbative contributions in the large-$`N`$ expansion of the gauge theory. In this way one could investigate to what extent these results apply to theories with less supersymmetry. It would, for instance, be interesting to find out how much can be said about the conformal field field theories described in \[17,,18\]. In such cases the dual string theory has an $`S^5/\mathrm{\Gamma }`$ sector which could admit world sheet instantons whose absence was one of the important ingredients in our argument. Acknowledgments: We would like to thank N. Berkovits, A. Lawrence, J. Maldacena, S. Minwalla, H. Osborn, A. Sen, A. Strominger and C. Vafa for useful discussions. We would also like to acknowledge the stimulating environment of the Cargese’99 workshop on Strings and M-theory as well as the Strings ’99 conference at Potsdam. R.G. would also like to thank the Abdus Salam International Centre for Theoretical Physics for it’s hospitality during the Extended Workshop on String Theory and Dualities, where part of this work was carried out. The research of R.G. is supported by DOE grant DE-FG02-91 ER40654. References relax J. Maldacena, The large $`N`$ limit of superconformal field theories and supergravity, hep-th/971120. relax S.S. Gubser, I.R. Klebanov and A.M. Polyakov, Gauge theory correlators from non-critical string theory, hep-th/9802109. relax E. Witten, Anti de Sitter Space and Holography, hep-th/9802150. relax H. Osborn (private communication); S. Gubser, I. Klebanov, Absorption by Branes and Schwinger Terms in the World Volume Theory, hep-th/9708042. Phys. Lett. 413B, 41 (1997); D. Anselmi, D.Z. Freedman, M.T. Grisaru, A.A. Johansen, Nonperturbative Formulas for Central Functions of Supersymmetric Gauge Theories, hep-th/970804; S. Lee, S. Minwalla, M. Rangamani, and N. Seiberg, Three-Point Functions of Chiral Operators in $`D=4`$ , $`𝒩=4`$ SYM at Large $`N`$, hep-th/9806074; E. D’Hoker, D.Z. Freedman, and W. Skiba, Field Theory Test for Correlators in the AdS/CFT Correspondence, hep-th/9807098; B. Eden, P.S. Howe, C. Schubert, E. Sckatchev, and P.C. West, Four-point Functions in $`𝒩=4`$ Supersymmetric Yang-Mills Theory at Two Loops, hep-th/9811172. relax T. Banks and M.B. Green, Nonperturbative effects in $`AdS_5\times S^5`$ string theory and d = 4 SUSY Yang-Mills, hep-th/9804170; JHEP 9805 (1998) 2. relax M. Bianchi, M.B. Green, S. Kovacs and G.C. Rossi, Instantons in supersymmetric Yang--Mills and D-instantons in IIB superstring theory, hep-th/9807033; JHEP 9808 (1998) 13. relax N. Dorey, T. J. Hollowood, V. V. Khoze, M. P. Mattis, S. Vandoren, Multi-Instanton Calculus and the AdS/CFT Correspondence in N=4 Superconformal Field Theory, hep-th/9901128. relax M.B. Green and M. Gutperle, Effects of D-instantons, hep-th/9701093, Nucl. Phys. B498 (1997) 195. relax J. Brodie, M. Gutperle, String corrections to four point functions in the AdS/CFT correspondence, hep-th/9809067, Phys. Lett. 445B,296 (1999). relax N. Berkovits and C. Vafa, Type IIB $`R^4H^{4g4}`$ Conjectures, hep-th/9803145; Nucl. Phys. B533 (1998). relax N. Berkovits, Generalization of the $`R^4`$ conjecture, Proceedings of Strings’98, http://www.itp.ucsb.edu/online/strings98/berkovits. relax J. G. Russo, Construction of SL(2,Z) Invariant Amplitudes in Type IIB Superstring Theory, hep-th/9802090; An Ansatz For A Nonperturbative Four Graviton Amplitude In Type IIB Superstring Theory, hep-th/9707241, Phys. Lett. B417 (1998) 253. relax J. Russo and A.A. Tseytlin, One-loop four-graviton amplitude in eleven-dimensional supergravity, hep-th/970713; Nucl. Phys. B508(1997)245. relax R. Gopakumar and C. Vafa, On the Gauge Theory/Geometry Correspondence, hep-th/9811131. relax G. ’t Hooft, “On the Convergence of Planar Diagram Expansions”, Commun.Math.Phys. 86, (1982), 449. relax N. Dorey, T.J. Hollowood, V.V. Khoze, M.P. Mattis and S. Vandoren, Yang-Mills instantons in the large N limit and the AdS / CFT correspondence, hep-th/9808157. relax S. Kachru, E. Silverstein, 4d Conformal Field Theories and Strings on Orbifolds, hep-th/9802183, Phys. Rev. Lett. 80, 4855, (1998). relax A. Lawrence, N. Nekrasov, C. Vafa, On Conformal Theories in Four Dimensions, hep-th/9803015, Nucl.Phys. B533, 199, (1998).
no-problem/9908/cond-mat9908331.html
ar5iv
text
# A model for the fragile-to-strong transition in water ## I Introduction The form of the dependence of viscosity on temperature is among the many properties that make water an anomalous fluid. Water is a fragile fluid when viewed at temperatures close to the melting temperature, indicating that there is an arrest of its degrees of freedom on cooling. This behavior is typical of many substance known as fragile glass-formers . However, close to the glass temperature $`T_G`$ ($`136`$ K)supercooled water shows characteristics of a strong liquid , in which there is an almost temperature independent configurational entropy, that manifests in an Arrhenius dependence of the viscosity $`\eta `$ as a function of $`T`$. Thermodynamic constraints limit the transition between these two regimes to occur rather sharply in a temperature range around $`220`$ K. There is by now a good piece of evidence that many of the anomalous properties of water can be rationalized by the use of an effective, two particle, spherical interaction potential, of the core-softened type. This interaction can be viewed as appearing between clusters of water molecules, rather than between single molecules. The main characteristic of this interaction is that it allows for two different equilibrium distances between clusters, depending on pressure. An appropriate, simplified model that capture many of the anomalies of water is provided by spherical particles interacting through a potential consisting of a hard core plus a soft repulsive shoulder . Here we show–using an analytically solvable version of it–that this kind of interaction can also explain the non-standard behavior of $`\eta (T)`$. ## II Hard spheres model We will use a model of hard spheres as a starting point (in the next section it will be generalized to describe the properties of water). We will suppose that the pure hard sphere system has an ideal thermodynamical glass transition at some temperature $`T_0`$ when the fluid phase is supercooled, preventing crystallization. A possible scenario for this glass transition is the following. For glassy systems there is a contribution $`s_c`$ to the entropy–referred to as configurational entropy–that comes from the many different configurations in which the glass can exist. For the case of spheres it comes from the many ways in which the spheres can be accommodated in stable, non-crystalline arrangements. These configurations differ in the value of the specific volume $`v`$. We will suppose that hard spheres have a configurational entropy per particle $`s_c^{HS}(v)`$ of the form<sup>*</sup><sup>*</sup>*In Ref. a parabolic form for $`s_c`$ as a function of density (instead of $`v`$) is used. The difference between both choices is tiny, and both give rise to an ideal glass transition. $$s_c^{HS}(v)=\alpha (vv_0)\beta (vv_0)^2.$$ (1) According to this formula, $`s_c^{HS}`$ becomes lower than zero for $`v<v_0`$ (and $`v>v_0+\alpha /\beta `$), indicating that there are no accessible states in this range, i.e., $`v_0`$ is the minimum value that $`v`$ can take. To get the total entropy $`s_{\mathrm{tot}}`$ of the system we should still include the contribution coming from small vibrations around each configuration . However, for our purpose this will not be necessary, since (as we will see below) $`\eta `$ is given in terms of $`s_c`$, rather than $`s_{\mathrm{tot}}`$, and $`s_c(T)`$ is independent of the vibrational contribution to the entropy. The previous form of the configurational entropy implies the existence of an ideal thermodynamical glass transition occurring at $`T_0`$, where $`T_0`$ is obtained from $`P/T_0=s_c^{HS}/v|_{v=v_0}=\alpha `$. Whereas from microscopic grounds there is no rigorous prove that this transition should occur, the consequences on observable magnitudes that can be predicted from it are consistent with the known phenomenology of glassy systems and with results of numerical simulations. For $`T<T_0`$ the system is in the fundamental configurational state, and then $`s_c^{HS}(T<T_0)=0`$. For $`T>T_0`$, $`s_c^{HS}`$ is given by $$s_c^{HS}(T>T_0)=\frac{\alpha ^2}{4\beta T^2}(T^2P^2/\alpha ^2)=\frac{\alpha ^2}{4\beta T^2}(T^2T_0^2)$$ (2) This expression for $`s_c^{HS}`$ can be used to calculate transport properties such as the viscosity $`\eta `$ through the use of the Adam-Gibbs formula . It states that the value of $`\eta `$ is given by $$\eta (T)=\eta _0\mathrm{exp}[A/(Ts_c)],$$ (3) which for hard spheres becomes $$\eta ^{HS}(T)=\eta _0\mathrm{exp}\left[4A\alpha ^2\beta T/(T^2T_0^2)\right]$$ (4) where $`\eta _0`$ and $`A`$ are constants. This is a behavior typical of a fragile glass-former, in which $`\eta `$ increases more rapidly than in a simple thermally activated process, and it diverges when $`TT_0`$. The presence of $`s_c`$ in (3) reflects the fact that jumps between different basins of the energy landscape become less probable as the number of these basins diminishes. ## III Core-softened models for water Properties of water have been studied recently by using models in which particles interact through potentials that allow for two different equilibrium distances between particles, namely $`d_0`$ and $`d_1>d_0`$. One possibility is, for instance, to take a strict hard core at $`d_0`$ and a shoulder that vanishes at $`d_1`$. Here we will use a further simplification of this kind of models in order to be able to extract analytical results. We consider spheres of radius $`r_1`$ ($`=d_1/2`$). Pairs of spheres will be allowed to overlap (more than two overlapping spheres will not be allowed), and each time this happens the system will be charged an energy $`\epsilon _0`$. This may be considered as a limiting case of particles with a core-softened potential, in which there is a low energy hard core at a distance $`2r_1`$, and a strict hard core at $`2r_0`$, and we are taking $`r_0=0`$. To make the problem analytically tractable, we will also suppose that each time two spheres overlap, they are constrained to have their centers in exactly the same position. This approximation neglects the entropy associated with small vibrations of the particles in each pair. We are interested in the configurational entropy $`s_c`$ of the system, now as a function of the specific enthalpy $`h=Pv+e`$, that includes the internal energy $`e`$ coming from the existence of overlapping particles. To calculate $`s_c(h)`$ we proceed in the following way. Suppose we have a system of $`N`$ particles, $`n`$ of them in non overlapped positions and $`n^{}`$ pairs of overlapped particles ($`N=n+2n^{}`$). The configurational entropy will be that of $`n+n^{}`$ hard spheres plus the combinatorial entropy for locating the $`n^{}`$ pairs in the $`n+n^{}`$ possible positions, i.e., $$\stackrel{~}{s}_c=\frac{n+n^{}}{N}s_c^{HS}+k_B\mathrm{ln}\left(\begin{array}{c}n+n^{}\\ n^{}\end{array}\right)$$ (5) (here we use $`\stackrel{~}{s}_c`$ to indicate an entropy functional). Using $`xn^{}/N`$ and $`\stackrel{~}{v}V/(n+n^{})=v/(1x)`$ as independent variables we can write $`\stackrel{~}{s}_c`$ as a $`\stackrel{~}{s}_c(x,\stackrel{~}{v})`$ $`=`$ $`(1x)s_c^{HS}(\stackrel{~}{v})+`$ (6) $`+`$ $`k_B\left[(12x)\mathrm{ln}\left({\displaystyle \frac{1x}{12x}}\right)+x\mathrm{ln}\left({\displaystyle \frac{1x}{x}}\right)\right].`$ (7) In order to get the thermodynamic value of $`s_c(h)`$ at each value of the external pressure $`P`$, we have to maximize $`\stackrel{~}{s}_c`$ for each fixed value of the enthalpy, i.e., $`s_c(h)`$ $`=`$ $`\underset{x,\stackrel{~}{v}}{\mathrm{min}}|_h\stackrel{~}{s}_c(x,\stackrel{~}{v})`$ (8) $`h=Pv+e`$ $`=`$ $`(1x)P\stackrel{~}{v}+x\epsilon _0.`$ (9) ## IV Results In Fig. 1 we see $`s_c`$, $`x`$, and $`v`$ as functions of $`h`$ for three different values of $`P`$, using expression (2) for $`s^{HS}(v)`$ with $`\alpha =2.79k_Br_1^3`$, $`\beta =0.97k_Br_1^6`$, and $`v_0=6.37r_1^3`$ which are values extracted from numerical simulations of hard spheres systems . We also plot in Fig. 1 the limiting cases $`\stackrel{~}{s}_c(x=0)`$ and $`\stackrel{~}{s}_c(x=0.5)`$, corresponding to all particles in singled or overlapped positions, respectively. The thermodynamic value $`s_c(h)`$ can never be lower than $`\stackrel{~}{s}_c(x=0)`$ or $`\stackrel{~}{s}_c(x=0.5)`$. The states with the lowest enthalpy for $`x=0`$ and $`x=0.5`$ have (from (9)) $`h=Pv_0`$ and $`h=Pv_0/2+\epsilon _0/2`$, respectively. These values coincide at $`P_{\mathrm{cr}}\epsilon _0/v_0`$. In Fig. 1(a), $`P=0.9P_{\mathrm{cr}}`$, $`\stackrel{~}{s}_c(x=0)`$ is always greater than $`\stackrel{~}{s}_c(x=0.5)`$ and for this reason $`x`$ takes values close to 0, indicating that most particles are in singled positions. The obtained $`s_c(h)`$ function departs from zero with infinite derivative at $`h=0.9\epsilon _0`$ (because of the combinatorial contribution to the entropy) but it still has a singularity (namely a jump in its second derivative) when $`s_c/h=\alpha /P`$. In (b) the value of $`P=1.1P_{\mathrm{cr}}`$ is larger, and at low $`h`$ the contributions with $`x=0.5`$ dominate, indicating that the system has almost all particles coupled in pairs. For higher $`h`$, $`x`$ goes down to zero, namely paired particles become rare. In (c), the value of $`P=P_{\mathrm{cr}}`$ is exactly that at which the ground state of the system with $`x=0`$ and the one with $`x=0.5`$ are degenerated. In this case, entropy starts from a finite value $`0.48k_B`$ at $`h_{\mathrm{min}}=\epsilon _0`$, corresponding to the maximum combinatorial entropy of choosing which particles are singled, an which are paired. To get the values of the thermodynamic variables as a function of $`T`$, instead of $`h`$, we have only to make use of the relation $`T^1=s/h`$. The results for the configurational entropy $`s_c(T)`$ and the viscosity $`\eta (T)`$ (calculated using the Adam-Gibbs formula (3)) are shown in Fig. 2. ## V Discussion and comparison with water From Fig. 2(a) we see that for any $`P`$, $`s_c`$ is finite for all $`T0`$, i.e., there is no vanishing of $`s_c`$ at any finite temperature, contrary to what happened in the case of simple hard spheres (see Eq. (2)). This is due to the possibility for the system of having particles singled or paired, which always accounts for the existence of a non-zero combinatorial entropy. For $`P=P_{\mathrm{cr}}`$ the bottoms of the $`\stackrel{~}{s}_c`$ functions corresponding to $`x=0`$ and $`x=0.5`$ coincide (Fig. 1(c)), and this combinatorial entropy can be used up to $`T=0`$, in such a way that $`s_c`$ remains finite, even when $`T0`$. For $`PP_{\mathrm{cr}}`$, $`s_c`$ goes to zero as $`T0`$, as the ground state is unique. There is still a phase transition at finite temperatures, signaled by the kink in the $`s_c(T)`$ curves in Fig 2(a). The position of this kink in the $`P`$-$`T`$ plane is given by $`P/T=\alpha `$. In the $`\mathrm{log}(\eta )`$-$`1/T`$ plot (Fig. 2(b)), the kinks mark the transition between a fragile and a strong behavior. At high $`T`$, $`s_c`$ diminishes rapidly with temperature and the system is fragile. This behavior is equivalent to that of the simple hard sphere system (see Eq. (4)). At low $`T`$, the dependence of $`s_c`$ on $`T`$ is much weaker, indicating a stronger behavior. In this regime, the available configurational entropy is mainly of combinatorial nature. For this model there is no ideal glass transition, i.e., $`\eta `$ is finite at any finite temperature, and $`T_0=0`$. However, from an experimental point of view, the glass transition $`T_G`$ is conventionally defined as the value of $`T`$ at which $`\eta `$ takes some large value (this is usually taken to be $`10^{13}`$ poise). In Fig. 3 we show curves of constant $`\eta `$ extracted from our model. Each of these may be thought as defining a dynamical glass transition temperature $`T_G`$ (that depends on pressure), according to different dynamical criteria. The position of the line of the fragile-to-strong transition is also indicated. We see that $`T_G`$ is systematically lower around the critical pressure $`P=P_{\mathrm{cr}}`$. This behavior has been observed in numerical simulations of SiO<sub>2</sub>, which has a fragile-to-strong transition qualitatively similar to that of water. Notice that $`T_G`$ is lower than the temperature of the fragile-to-strong transition only in some range of pressure around $`P_{\mathrm{cr}}`$, and only in this range the fragile-to-strong transition will be experimentally observable. One important ingredient that we have not included in the model is the existence of an attractive part in the interaction potential. In real water this attraction generates the liquid-vapor first order coexistence line, and also probably a second first order line in the supercooled region separating two different amorphous configurations. A simple way of analyzing the consequences on our model of an attraction between particles is the following. If the attraction is considered to be long ranged, of van der Waals type, then all the results we have obtained remain valid if we replace $`P`$ by a new effective pressure $`P^{}P+\gamma /v^2`$ with some constant $`\gamma `$, namely, the attraction acts as an effective pressure (which depends on $`v`$) that has to be added to the external pressure $`P`$. In the $`P`$-$`T`$ phase diagram, this non-uniform transformation of the $`P`$ axis produces (if $`v`$ decreases rapidly when $`P`$ increases) a “folding” that indicates a first order transition. This is the way in which the liquid-vapor coexistence line appears in the van der Waals equation for a fluid. For our model, since $`v(P)`$ at $`T=0`$ has an abrupt discontinuity at $`P=P_{\mathrm{cr}}`$, the attraction generates also a new first order line ending in a critical point. It seems to be that in water this critical point lies at temperatures higher than $`T_G`$, so the first order line determines two different sectors of the glassy phase of water. These two sectors correspond to the experimentally observed high density amorphous (above the first-order line) and low density amorphous (below the first order line) phases. In our model, these two phases differ in the fraction of particles that are paired, and thus they can also be named high density and low density amorphous phases. We have relied for our discussion upon the existence of a thermodynamic phase transition for hard spheres, that is not rigorously proven to occur. However, for slightly different forms for $`s_c^{HS}`$ (which may imply the absence of an ideal glass transition) than that given by Eq. (2), our results still remain valid, except for the fact that the sharp fragile-to-strong transition (the kinks in the curves of Fig. 2) becomes a crossover. The model we have presented explains the transition between fragile and strong behavior of water as appearing from the competition between two different local structures. To be able to solve the problem analytically, we had to make the crude approximation that these structures correspond to singled and paired particles. In real water it is likely that what plays the role of our particles are the so-called Walrafen pentamers, which are clusters of five water molecules. These cluster are naturally expected to accommodate at one of two possible distances from each other. It is clear that in this more general case the mechanism for a fragile-to-strong transition may remain basically the same. In fact, from the way we solved the model, it is seen that all what is needed is the existence of two different competing structures, independently of the details of them. At high $`T`$ the fragile behavior is associated to the configurational entropy of each structure individually. At low $`T`$ the strong behavior appears due to the combinatorial entropy of choosing locally between the two structures.
no-problem/9908/hep-th9908173.html
ar5iv
text
# Loss of Causality in Discretized Light-Cone Quantisation ## Introduction Microcausality is a fundamental postulate in relativistic quantum field theory. Mathematically, it states that two local operators $`𝒪_1(x)`$ and $`𝒪_2(y)`$ at space-time positions $`x`$ and $`y`$ must commute if $`xy`$ is a space-like distance. Physically, it means that signals cannot be transmitted faster than with the velocity of light, $`c`$. Imposing the requirement of microcausality eliminates a large class of possible quantisation schemes. For instance, quantising scalar fields with anti-commutators violates microcausality, enforcing quantisation of scalars in terms of commutators. Along these lines, the celebrated spin-statistics theorem is established. In this letter we demonstrate that the popular method of front form (FF) or light-cone (LC) quantisation leads to a breakdown of microcausality when space-time is compactified to a cylinder with the periodic direction being chosen as light-like. As the momenta conjugate to this direction become discrete, the method is usually referred to as discretized light-cone quantisation (DLCQ) . Its range of applicability has recently been extended to include M-theory . We will be concerned with a massive scalar field in $`d`$ space time dimensions. After some general remarks we will actually specialise to $`d`$ =2. Our notations and conventions are $`x^\pm =x^0\pm x^{d1},k^\pm =k^0\pm k^{d1},`$ (1) $`xk=g_{\mu \nu }x^\mu k^\nu ,g_+=g_+=1/2.`$ (2) FF quantisation (for a recent review see ) amounts to prescribing field commutators on the quantisation surface $`x^+=0`$. This is a hyperplane tangent to the LC which in $`d=2`$ collapses to a light ray. It has been shown by a number of authors , that quantisation on just one light-like surface is ambiguous. Knowledge of initial conditions on two quantisation surfaces, say $`x^+=0`$ and $`x^{}=0`$, is necessary in order to have a well-posed (characteristic) initial-value problem . As a result, the characteristic initial values uniquely determine solutions to the Klein-Gordon equation for all $`x^\pm `$ larger than the initial ones. In 1994, however, Heinzl and Werner were able to show that the introduction of periodic boundary conditions (pBC) in $`x^{}`$ direction (assumed to be compact) uniquely determines the fields on the second quantisation surface, $`x^{}=0`$, in terms of the fields on the first quantisation surface, $`x^+=0`$. An infinite-volume formulation of this problem, however, is rather involved and in general will require the use of distribution theory . It will not be addressed in this letter. We will rather restrict to the finite-volume case and show that prescribing light-like BC—though solving the initial-value problem—is in conflict with microcausality. The results presented here are built upon the Ph.D. thesis where the loss of causality in DLCQ has been reported for the first time. Recently, this finding has been confirmed by other researchers . ## Violation of Microcausality A. Generalities We start with a free scalar field $`\varphi `$ in $`d`$ space-time dimension. The commutator of two free scalar fields is of course known for all times (i.e. everywhere in Minkowski space), $`[\varphi (x),\varphi (0)]=i\mathrm{\Delta }(x)`$. $`\mathrm{\Delta }(x)`$ denotes the Pauli-Jordan or Schwinger function, $$\mathrm{\Delta }(x)=\frac{1}{i}\frac{d^dk}{(2\pi )^{d1}}\delta (k^2m^2)\text{sgn}(k^0)e^{ikx}.$$ (3) As the sign of $`k^0`$ does not change under proper orthochronous Lorentz transformations, $`\mathrm{\Lambda }_+^{}`$, if $`k^2`$ is spacelike, $`\mathrm{\Delta }`$ is a Lorentz invariant function. The sign function in addition guarantees that $`\mathrm{\Delta }`$ is antisymmetric, $`\mathrm{\Delta }(x)=\mathrm{\Delta }(x)`$, as is necessary for a commutator. It is well known that $`\mathrm{\Delta }(x)`$ obeys microcausality. This is the statement that $`\mathrm{\Delta }(x)`$ has to vanish for $`x`$ space-like, i.e. $`x^2<0`$. A very elegant argument to see this is due to Gasiorowicz . For $`x^2<0`$, there is a Lorentz transformation $`\mathrm{\Lambda }_+^{}`$ that takes $`x`$ to $`x`$, thus, by invariance, $`\mathrm{\Delta }(x)=\mathrm{\Delta }(x)`$, for $`x^2<0`$. Therefore, outside the LC, $`\mathrm{\Delta }`$ is both symmetric and antisymmetric in $`x`$ and must vanish. The argument does not work for $`d=2`$, as the regions $`x^1>0`$ and $`x^1<0`$ are disconnected. Nevertheless, causality also holds in $`d=2`$, as one can see upon evaluating (3) for this case, $`\mathrm{\Delta }(x)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\text{sgn}(x^0)\theta (x^2)J_0(m\sqrt{x^2}),`$ (4) $`=`$ $`{\displaystyle \frac{1}{4}}\left[\text{sgn}(x^+)+\text{sgn}(x^{})\right]J_0(m\sqrt{x^2}),`$ (5) which indeed vanishes outside the LC (here $`J_0`$ denotes the Bessel function). The restrictions of (4) to $`x^0=0`$ and $`x^+=0`$ yield the canonical commutators of the two quantisation schemes . In (4) we have given both the IF and FF versions of $`\mathrm{\Delta }`$ which are, of course, simply related by the coordinate transformation (1). These two forms can actually be represented as one dimensional integrals by performing the energy integrations over $`k^0`$ and $`k^{}`$, respectively, $`\text{IF:}\mathrm{\Delta }(x)`$ $`=`$ $`{\displaystyle \frac{dk^1}{2\pi \omega _k}\mathrm{sin}(\omega _kx^0k^1x^1)},`$ (6) $`\text{FF:}\mathrm{\Delta }(x)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk^+}{2\pi k^+}}\mathrm{sin}(\widehat{k}^{}x^+/2+k^+x^{}/2).`$ (7) The on-shell values of the energies are given by $`\omega _k=(k_1^2+m^2)^{1/2}`$ and $`\widehat{k}^{}=m^2/k^+`$. Note the restriction of the integration in (7) which is due to the positivity of the longitudinal momentum $`k^+`$. Both representations (6) and (7) can be integrated and yield (4). As a cross check we note that (6) and (7) are still related by the coordinate transformation (1) applied to the *on-shell* momenta, $$k^\pm =\omega _k\pm k^1=\sqrt{k_1^2+m^2}\pm k^1.$$ (8) This makes the positivity of $`k^+`$ explicit and entails that the integration measures are related by the singular transformation $`dk^1/\omega _k=dk^+/k^+`$. Let us now investigate how (6) and (7) get modified in a finite volume. To this end we restrict the spatial coordinates, $`Lx^1,x^{}L`$, and impose pBC for the field $`\varphi `$. The conjugate momenta become discrete, $`k_n^1\pi n/L`$ and $`k_n^+2\pi n/L`$, respectively. The finite volume representations are defined by replacing the integrals (6) and (7) by the discrete sums, $`\mathrm{\Delta }_{IF}(x)`$ $``$ $`{\displaystyle \underset{n=N}{\overset{N}{}}}{\displaystyle \frac{1}{2\omega _nL}}\mathrm{sin}(\omega _nx^0n\pi x^1/L),`$ (9) $`\mathrm{\Delta }_{FF}(x)`$ $``$ $`{\displaystyle \underset{n=1}{\overset{N}{}}}{\displaystyle \frac{1}{2\pi n}}\mathrm{sin}(\widehat{k}_n^{}x^+/2+n\pi x^{}/L),`$ (10) where the limit $`N\mathrm{}`$ is understood. The on-shell energies for discrete momenta are defined as $`\omega _n=(n^2\pi ^2/L^2+m^2)^{1/2}`$ and $`\widehat{k}_n^{}=m^2L/2\pi n`$. For both functions, $`\mathrm{\Delta }_{IF}`$ and $`\mathrm{\Delta }_{FF}`$, the periodicity in $`x^1`$ and $`x^{}`$, respectively, with periodicity length $`2L`$, is obvious. Note that $`\mathrm{\Delta }_{IF}`$ contains a zero mode ($`n`$ = 0), while $`\mathrm{\Delta }_{FF}`$ does not. This is a consequence of the Klein-Gordon equation . In what follows we will show that, unlike $`\mathrm{\Delta }_{IF}`$, $`\mathrm{\Delta }_{FF}`$ does not obey microcausality, i.e. does not vanish for $`x^2x^+x^{}<0`$. In addition, we find that $`\mathrm{\Delta }_{FF}`$ does not converge to $`\mathrm{\Delta }`$ in the infinite volume limit. It actually turns out that BC in a light-like direction are quite generally incompatible with causality. B. Numerical Results In general, the sums (9) and (10) cannot be evaluated analytically. Therefore we have calculated them numerically. The summation cutoff $`N`$ has been chosen sufficiently large to establish numerical convergence. The results are shown in Fig. 1 and Fig. 2. By comparing the two figures one observes a striking difference. $`\mathrm{\Delta }_{IF}`$ is a smooth and regular function, while $`\mathrm{\Delta }_{FF}`$ looks ‘noisy’ and irregular. Furthermore, for fixed $`0<x^0<L`$, $`\mathrm{\Delta }_{IF}`$ has compact support *inside* the LC, $`x^0<x^1<x^0`$ (and in the periodic copies of this interval). Outside the LC $`\mathrm{\Delta }_{IF}`$ shows tiny oscillations around the value zero, which vanish in the limit $`N\mathrm{}`$. The oscillations are due to Gibbs’ phenomenon (the Fourier series does not converge uniformly in the vicinity of points where the limiting function makes jumps). Physically, what happens is that we have point sources located at positions $`x^1=2Ln`$. These ‘emit’ spherical waves which do not interfere unless $`x^0L`$. For $`x^0>L`$ (not shown here) we have an interference phenomenon so that $`\mathrm{\Delta }_{IF}`$ no longer vanishes outside the LC, which is a straightforward consequence of periodicity. The situation concerning $`\mathrm{\Delta }_{FF}`$ is different. Numerically, one sees that despite the irregular shape the sum (10) converges to a periodic function. The most important observation, however, is that $`\mathrm{\Delta }_{FF}`$ does not vanish outside the LC, i.e. for $`x^{}<0`$, $`x^+>0`$ as in Fig. 2. This a clear *violation of microcausality*. We have numerical evidence for a corresponding behavior in $`d=3,4`$ space-time dimensions. C. Analytical Results Let us try to get an analytical understanding of the numerical results beginning with $`\mathrm{\Delta }_{IF}`$. A straightforward application of the Poisson resummation formula yields $$\mathrm{\Delta }_{IF}(x)=\underset{n}{}\mathrm{\Delta }(x^0,x^1+2Ln),$$ (11) with the continuum $`\mathrm{\Delta }`$ from (4). This result is exactly what we see in Fig. 1, a periodic array of (nearly) smooth functions with support inside the LC (and its periodic copies). It should be stressed that $`\mathrm{\Delta }_{IF}`$ is causal even for *finite* extension $`L`$, i.e. without the infinite-volume limit being performed. Let us now analyze $`\mathrm{\Delta }_{FF}`$. First note the rather weak localization properties of $`\mathrm{\Delta }`$ in the LC direction $`x^{}`$. For positive LC time $`x^+`$, $`\mathrm{\Delta }`$ vanishes outside the LC, i.e. for $`x^{}<0`$, and decays slowly for $`x^{}>0`$, asymptotically like $`(x^{})^{1/4}`$. The integrand in (7), denoted by $`I(k^+)`$, oscillates rapidly for small $`k^+`$ such that the zero mode, i.e. the limit $`I(k^+=0)`$, is not defined. It turns out that this makes the application of Poisson resummation impossible. Because the latter cannot be used, let us consider the following alternative which leads to an analytic and close approximation of $`\mathrm{\Delta }_{FF}`$. For this purpose we rewrite (10) in terms of dimensionless variables (cf. Fig. 2), $$\mathrm{\Delta }_{FF}(v,w)=\underset{n=1}{\overset{N}{}}\frac{1}{2\pi n}\mathrm{sin}(w/2\pi n+2\pi nv),$$ (12) with $`vx^{}/2L`$, $`wm^2L^2(x^+/2L)`$. If we expand (12) in powers of $`w`$, the sum over $`n`$ can actually be performed using summation formulae 1.443.1/2 from . With the restriction $`1v1`$, the result is $$\mathrm{\Delta }_{FF}(v,w)=\frac{1}{2}\underset{n=0}{\overset{N}{}}\frac{w^n}{n!(n+1)!}\text{sgn}^{(n+1)}(v)B_{n+1}(|v|),$$ (13) where $`B_n`$ denotes the $`n`$th Bernoulli polynomial . The series (13), being a power series instead of a Fourier series, converges rather rapidly as a function of $`w`$. In addition, the limit is approached uniformly (no Gibbs phenomenon). This is obvious from Fig. 3 where we compare the resummed expression (13) with the Fourier representation (12) (for $`w`$ = 5). The agreement is quite impressive which is to be expected as we have summed the first 20 terms in (13). Numerically, one finds convergence if the number $`N`$ of terms summed over is of the order of $`w/2`$. This is due to the fact that the amplitude of the Bernoulli polynomials $`B_n`$ decreases rapidly with $`n`$. Thus, for $`w2`$, i.e. $`x^+4/m^2L`$, the first two terms in the expansion (13) are already a rather good approximation so that we can write, $$\mathrm{\Delta }_{FF}(v,w)\frac{1}{4}\text{sgn}(v)+\frac{v}{2}\frac{w}{4}(v^2|v|+1/6).$$ (14) This result provides an *analytical* check that $`\mathrm{\Delta }_{FF}`$ does not vanish outside the LC ($`1<v<0`$), and thus, that causality is violated. ## Restoration of a Causal Propagator With representation (10) we are sampling a continuous function $`I(k^+)`$ by *equidistant* points on a momentum grid in $`k^+`$. For small $`k^+`$, however, this is a very bad approximation, as $`I(k^+)`$ is rapidly oscillating there, with a frequency increasing roughly as $`1/k^+`$. The point $`k^+=0`$ is thus an accumulation point of the Fourier spectrum. In its vicinity, we should actually sample with a momentum resolution $`\mathrm{}k_n^+1/n`$. In other words, instead of harmonic one should use *anharmonic* “Fourier” analysis. If we use (8) to introduce *new* discrete longitudinal momenta, $$k_n^\pm \omega _n\pm k_n^1=\frac{1}{L}(\sqrt{n^2\pi ^2+m^2L^2}\pm n\pi ),$$ (15) we can write down the causal, finite-volume commutator in terms of light-cone variables, $$\mathrm{\Delta }_c(x)=\frac{1}{L}\underset{n}{}\frac{k_n^+}{(k_n^+)^2+m^2}\mathrm{sin}(k_nx),$$ (16) where $`k_nx=k_n^{}x^+/2+k_n^+x^{}/2`$. Obviously, the momentum grid (15) used in (16) is not equidistant. In particular, for small $`k_n^+`$, corresponding to large negative $`k_n^1`$, one finds $`\mathrm{}k_n^+1/n`$. Thus, the small-$`k^+`$ region becomes sampled in a reasonable way such that the features of $`I(k^+)`$, which guarantee the causality of $`\mathrm{\Delta }`$ are properly described even on a finite (momentum) lattice (see Fig. 4). ## Discussion The above analysis shows that one cannot have both, periodicity in $`x^{}`$ and causality of the commutator $`\mathrm{\Delta }(x^+,x^{})`$. If one insists on periodicity, one violates causality and vice versa. Consequently, the method of DLCQ in which the field operators are expanded in periodic plane waves, is non-causal. This does not come as too big a surprise: it is well known that DLCQ yields a rather poor representation of the small-$`k^+`$ behaviour of observables . Using the relation between Pauli-Jordan function and Feynman propagator, $`\pi \mathrm{\Delta }(k)=\text{sgn}(k^+)\text{Im}\mathrm{\Delta }_F(k)`$, we see that a causality violation also affects the Feynman propagator $`\mathrm{\Delta }_F`$. Note that the causality of $`\mathrm{\Delta }`$ can be viewed as a delicate cancellation between particle and anti-particle propagation amplitudes. Therefore it seems that also charge conjugation symmetry is violated by imposing light-like periodicity. Furthermore, with microcausality being at the heart of any dispersion relation, one expects problems also there. We have seen that one can taylor an ad hoc momentum grid with a special anharmonic resolution which remedies the causality violation of the *commutator*. It is an open question, however, whether this solves the causality problem of DLCQ in general. We expect the answer to be negative: any causal Green function will have its own peculiar small-$`k^+`$-behaviour and thus will require its own momentum grid which generically will be different from the one introduced above. An ensuing dependence of results on a particular discretisation choice clearly cannot be accepted. Acknowledgements T.H. thanks E. Elizalde, F. Lenz, B. Schroer, T. Tok, and A. Wipf for valuable hints and discussions. The authors gratefully acknowledge support by DFG, project WI 777/3-2 (T.H.), NSERC Canada (H.K.) and Austrian Science Fund, project P111098 PHY (N.S.).
no-problem/9908/cond-mat9908264.html
ar5iv
text
# Self-Organized Criticality in the Olami-Feder-Christensen model ## Abstract A system is in a self-organized critical state if the distribution of some measured events obeys a power law. The finite-size scaling of this distribution with the lattice size is usually enough to assume that the system displays SOC. This approach, however, can be misleading. In this work we analyze the behavior of the branching rate $`\sigma `$ of the events to establish whether a system is in a critical state. We apply this method to the Olami-Feder-Christensen model to obtain evidences that, in contrast to previous results, the model is critical in the conservative regime only. PACs number(s): 64.60.L, 05.40, 05.70.L Keywords: SOC, Random Processes, and Non-equilibrium Thermodynamics. In spite of many efforts and more than a decade of studies, the presence of self-organized critical behavior in nature (and in some computer models) is a matter of controversy. The concept of self-organized criticality (SOC) was originally proposed by Bak, Tang and Wiesenfeld to describe the appearance of scale invariance in nature. The idea was presented through the study of the behavior of avalanches in a sandpile ‘toy’ model . This simple model displayed the fundamental properties associated with self-organized criticality. Under a slow driven perturbation the system evolves to a critical state, with no characteristic time and length scales. Once in this state, the response of the system to the slow perturbation has no typical length, and even a small perturbation (as the addition of a single grain of sand) can start a big avalanche. Avalanching behavior as well as scale invariance have been experimentally observed in a variety of situations in nature, ranging from such different phenomena as earthquakes or magnetic systems (the Barkhausen effect), to biological problems such as evolution of species or lung inflation, just to give some examples. Although by now the initial attempt to explain the appearance of all linear scaling in nature through the concept of SOC may seem a little naive, the ubiquity of its presence is still a strong suggestion that some kind of ‘robust’ and general mechanism may be behind many of these phenomena. The concept of SOC has become polemic, and, up to now, there is no general agreement about the ingredients necessary to create the self-organized critical state. Particularly, there are discussions about the need of some kind of local conservation as an essential ingredient of the system to display SOC. The existence of SOC in non-conservative models would be highly desirable in this context, since, in practice, some kind of dissipation is always present in nature. One of the best successful applications of the ideas of SOC for non-conservative systems are the investigations of the Olami-Feder-Christensen on a model for the dynamic of earthquakes (hereafter called OFC model). In this model there is a parameter $`\alpha `$ that controls the level of conservation. Based on strong numerical evidences it has been taken as an example of a system that has self-organized criticality in the non-conservative regime, that is, for $`\alpha <0.25`$. In this paper we revisited the OFC model, but with a different technique. Instead of looking for power laws in distribution functions of avalanche sizes versus lattice sizes, we looked at the behavior of the average branching rate, both in the conservative and in the non-conservative regime. In contrast to previous evidences, we concluded that the OFC model is critical only for $`\alpha 0.25`$ (that is, in the conservative regime). For values of $`\alpha `$ close to but smaller than $`0.25,`$ this model could be classified as ‘almost critical’. That means that, although being critical only for $`\alpha =0.25`$, for all practical purposes the system behaves as if it were critical for a wide range of values of $`\alpha `$ , with well-defined power laws over many decades. In a recent paper, Kinouchi and Prado showed that some models that exhibit numerical evidences of self-organized criticality in a wide range of the coupling parameters were indeed what they called ‘almost critical’. Through the analysis of the branching rate $`\sigma `$ as a function of the dissipation parameter $`\alpha ,`$ they have shown that, although those systems are critical only for $`\alpha =\alpha _c`$, there are a rather large region around this point where approximate scale invariance holds. They called this behavior ‘almost critical’ since, in practice, it can hardly be distinguished from ‘true’ criticality based on the usual numerical evidences only. By usual numerical evidences we mean power-law behavior and scale invariance in distribution functions (the avalanche size distribution function, for instance). They also suggested that the analysis of the branching rate $`\sigma `$ (where $`0\sigma 1`$) as a function of the coupling constant $`\alpha `$ could be a more efficient way to determine whether a model is critical or not. To look for power-laws in lattices of increasing sizes is not a very efficient way to determine if a system is in fact critical, and this approach has already lead to mistakes. If the analyzed lattices are not big enough, the distribution functions of avalanche sizes $`F(s)`$ are power laws, even if the model does not display SOC. Because the computational cost of simulating the OFC model (and many others) in big lattices is prohibitive and there is no way to know, beforehand, if the considered lattices are big enough to show the real characteristics of the dynamical behavior of the system, such approach is hardly conclusive. It has been shown that some SOC models, with no spatial correlations and in the thermodynamic limit, can be mapped into a branching process. A branching process is a Markovian process and can be characterized by a sequence of random variables $`\left\{P(n)\right\}_{n=0}^{\mathrm{}}`$ , $`nN`$, where $`P(n)`$ represents the total number of individuals of the $`n^{th}`$ generation. Consider a group of individuals (ancestors) that can replicate, giving birth to some descendants, and let $`p_i`$ $`(i=0,1,\mathrm{},\mathrm{})`$ be the probability of an ancestor to give birth to $`i`$ descendants. Each of its descendant in turn can give birth to other descendants with the same probability $`p_i`$ so that $`p_i`$ does not depend on the previous generations and on the number of descendants of other individuals in the same generation. The branching rate, $`\sigma =\underset{i=0}{\overset{\mathrm{}}{}}ip_i`$, is then defined as the average number of descendants per ancestor. It is a well known result that, in order to have a critical branching process, one must have $`\sigma =1`$. Then the total number of descendants $`P(n)`$ in each generation (the size of the ‘colony’) behaves as a power law $`P(n)n^{3/2}`$ . On the basis of these considerations about the branching rate, and using different approaches, several authors were able to show that the random version of the OFC model was critical in the conservative regime only. Therefore, we decided to use this same approach to revisit the original Olami-Feder-Christensen model. This coupled-map lattice model is inspired on the spring block model developed by Burridge and Knopoff . Each site $`(i,j)`$ of a square lattice is associated with a continuous ‘energy’ $`F_{ij}`$, initially set to a random value in the interval $`(0,F_c)`$. The system is driven by a global perturbation that increases the energy of all sites uniformly and simultaneously. This process goes on until eventually one site becomes supercritical, that is, $`F_{ij}F_c`$ . This site becomes unstable and the system then relaxes according to the rules $`F_{ij}`$ $``$ $`0,`$ and $`F_{nn}`$ $``$ $`F_{nn}+\alpha F_{ij},`$ where $`F_{nn}`$ are the four nearest neighbors of site $`(i,j)`$. The parameter $`\alpha `$ controls the level of conservation of the model. If $`\alpha =0.25`$, the system is said to be ‘conservative’, that is, all the energy (or strength) lost by the site $`(i,j)`$ is distributed to its neighbors. This relaxation rule can possibly produce a chain reaction that only ends when all sites are stable again $`(F_{ij}<F_c,i,j)`$. As in the original work, we assume open boundaries. Also, as shown in reference, one must have $`\alpha <0.25`$ to mimic the dynamic of a real earthquake (some ‘energy’ or ‘strength’ is always lost to the upper moving tectonic plate). This model is believed to display self-organized criticality even when the dynamic is non-conservative $`(0<\alpha <0.25)`$. This is a result not yet fully understood, and it has been a matter of controversy the value of the lower bound for $`\alpha `$ (if it exists), under which the system has a localized behavior (note that we know that $`\alpha =0\sigma =0,`$ and, for $`\alpha =0.25,`$ we should have $`\sigma =1`$). Because it is a model defined on a lattice, analytical approaches are difficult and most of the results have been obtained from computer simulations. As the existence of a lattice introduces spatial correlations, it is not possible to define the probability $`p_i`$ analytically. We estimate the branching rate $`\sigma `$ numerically ($`\sigma =<n_d>`$, where $`<n_d>`$ is the average number of supercritical sites (descendants) originated by an unstable site). Just for comparison, we also study the random neighbor version of the OFC model (R-OFC), for which there are some analytical results showing that the model is critical for $`\alpha =0.25`$ only. Our results are presented in Tables 1 and 2 and in Figures 1 to 3. We checked the dependence of $`\sigma `$ on the lattice size (see Figures 2 and 3), and a special care has been taken to guarantee that the long transients were eliminated. We also checked the effects of the boundaries. In the OFC model we considered open boundaries to calculate $`\sigma ,`$ taking into account that the average number of descendants for a boundary site is the number of unstable sites it gives birth divided by the real number of neighbors of the ‘ancestor’ site (3 for a border site and 2 for a corner site). The R-OFC model was simulated without borders. In most of the cases, we first generated different stationary configurations from different random initial configurations. The errors were estimated by averaging results obtained for different initial configurations of the lattice (the errors so obtained are usually bigger than the ones obtained by averaging $`\sigma `$ during many generations, except when the system is conservative). The number of iterations needed to reach the stationary state is very big, and grows with the lattice size. In the OFC model the transient is bigger for smaller values of $`\alpha `$, while in the R-OFC the transient grows as $`\alpha `$ grows, making it impossible to simulate the case $`\alpha =0.25`$ (the point in the graph in this case was obtained from theoretical results). Once we were sure to have a stationary configuration, we analyzed the statistics of $`\mathrm{100\hspace{0.17em}000}`$ to $`\mathrm{5\hspace{0.17em}000\hspace{0.17em}000}`$ avalanches in the stationary state, to obtain (a) the average avalanche size $`<s>`$, (b) the branching rate $`\sigma `$ (weighting border sites), (c) the branching rate in the bulk $`\sigma _b`$ (taking into account only sites in the bulk), and (d) the average number of generations in an avalanching process $`<n>`$. Table 1 shows the results for the OFC model and Table 2 shows the results for the R-OFC model. There are no relevant differences between the behaviors of the OFC and the R-OFC models. For both of them, $`\sigma (\alpha )1`$ smoothly from below as $`\alpha 0.25`$ , with no sign of any kind of discontinuity in its behavior. Also, as can be seen in Figure 1, $`\sigma _{OFC}<\sigma _{ROFC},`$ for $`0.22\alpha <0.25`$. From theoretical considerations, we know that $`\sigma _{ROFC}<1`$ for $`\alpha <0.25`$. In Figures 2 and 3 we present the dependence of $`\sigma `$ on the lattice size for the OFC and the R-OFC models. These figures show that $`\sigma `$ grows almost linearly with $`1/L`$ with no suggestion that $`\sigma 1`$ as $`1/L0`$. The behavior of the system seems to be qualitatively different only if $`\alpha =0.25`$ (conservative case). We also checked the dependence of $`\sigma `$ on the generation $`n`$ within an avalanching process. We see that $`\sigma (n)`$ converges relatively fast to an asymptotic value . None of our conclusions were affected if we considered these asymptotic values of $`\sigma (n)`$ instead of the average value. The existence of SOC in the non-conservative regime of the OFC model has been accepted based mainly on numerical results of a work done by Middleton and Tang in 1995. In this paper, the authors showed how the natural tendency of this model to synchronize is destroyed by inhomogeneities introduced by the asymmetries of the boundaries, creating long-range correlations and leading to a power-law behavior in the distribution of avalanche sizes. The apparent contradiction between this result and ours can be understood from the conclusions of Kinouchi and Prado . In this paper, the study of two different models with an analytical solution (the extremal Feder and Feder model, EFF, with and without noise), shows that the effect of noise is to enlarge the region where the system displays an apparent critical behavior, leading to what was called ‘almost criticality’. The EFF model with noise displays a power law behavior (although it is not critical). In contrast, in the noiseless model, large avalanches occur in the conservative limit only. This also seems to be the case of the OFC model. The randomness introduced by the asymmetries of the boundaries creates correlations that enlarge the critical region leading to an ‘almost critical’ behavior, although it is not enough to ensure true criticality. In conclusion, we showed that the analysis of $`\sigma (\alpha )`$ is a complementary approach to define if a model is or is not critical. This new method revealed that the behavior of the OFC model is qualitatively identical to the behavior of the R-OFC. In contrast to previous results, the Olami-Feder-Christensen model seems to be critical only in the conservative regime, that is for $`\alpha =0.25`$. Both models are ‘almost’ critical in the sense defined in reference : $`\sigma 1`$ when $`\alpha 0.25`$, leading to a power law behavior of the avalanche sizes for many decades, and making it (almost) impossible to distinguish this behavior from ‘true’ self-organized criticality based on the observation of power-laws and finite-size scaling fits. The authors acknowledge Dr. Osame Kinouchi for helpful discussions and suggestions. J. X. Carvalho acknowledges the Brazilian agency CAPES for financial support. FIGURE CAPTIONS Table 1:Values of $`\sigma `$, $`\sigma _b`$ and $`<s>`$ for different values of the conservative parameter $`\alpha `$ in the Olami-Feder-Christensen (OFC) model. Results presented are those obtained with the biggest lattice ($`L_{max}`$) we were able to simulate. They represent the average of results obtained for different initial configurations and the errors are the errors associated with those averages. Table 2:Values of $`\sigma `$, $`\sigma _b`$ and $`<s>`$ for different values of the conservative parameter $`\alpha `$ in the random version of the Olami-Feder-Christensen (R-OFC) model. Results presented are those obtained with the biggest lattice ($`L_{max}`$) we were able to simulate. They represent the average of results obtained for different initial configurations and the errors are the errors associated with those averages. Figure 1: Branching rate as a function of the conservation parameter $`\alpha `$. Squares refer to the Olami-Feder-Christensen model (OFC) and circles to the Random version of the OFC model (R-OFC). In all cases the lattice size is $`L=100.`$ Figure 2: Branching rate as a function of the inverse of lattice size $`(1/L)`$ for the Random version of Olami-Feder-Christensen model. Different curves refer to different levels of conservation ($`\alpha =0.22,\mathrm{\hspace{0.17em}0.23}`$ and $`0.24`$). We can see that even for $`\alpha =0.24,`$ if we let $`L\mathrm{}`$, the branching rate $`\sigma `$ tends to a value smaller than $`1.`$ The system shows a qualitatively different behavior only if $`\alpha =0.25`$. Figure 3: Branching rate as a function of the inverse of lattice size $`(1/L)`$ for the Olami-Feder-Christensen model. Different curves refer to different levels of conservation ($`\alpha =0.23,\mathrm{\hspace{0.17em}0.24}`$ and $`0.25`$). We can see that even for $`\alpha =0.24,`$ if we let $`L\mathrm{}`$, the branching rate $`\sigma `$ tends to a value smaller than $`1.`$ Note that in the conservative case ($`\alpha =0.25`$) $`\sigma `$ is almost 1.00 even to very small lattices.
no-problem/9908/cond-mat9908076.html
ar5iv
text
# Coulomb blockade in low mobility nanometer size Si:MOSFETs \[ ## Abstract We investigate coherent transport in Si:MOSFETs with nominal gate lengths 50 to 100nm and various widths at very low temperature. Independent of the geometry, localized states appear when $`Ge^2/h`$ and transport is dominated by resonant tunnelling through a single quantum dot formed by an impurity potential. We find that the typical size of the relevant impurity quantum dot is comparable to the channel length and that the periodicity of the observed Coulomb blockade oscillations is roughly inversely proportional to the channel length. The spectrum of resonances and the nonlinear I-V curves allow to measure the charging energy and the mean level energy spacing for electrons in the localized state. Furthermore, we find that in the dielectric regime the variance $`var(lng)`$ of the logarithmic conductance $`lng`$ is proportional to its average value $`<lng>`$ consistent with one-electron scaling models. \] After the pioneering work of Scott-Thomas et al , Coulomb blockade in quantum dots formed by an impurity potential has been studied in quasi 1D wires or point-contact geometries . In comparison with lithographically defined lateral quantum dots, impurity quantum dots (IQD) contain typically fewer electrons. Downscaling the size of the IQD allows to operate a silicon based single electron quantum dot transistor even at room temperature . In the opposite case of wires, i.e. in disordered thin and wide insulating barriers, no Coulomb blockade oscillations have been reported up to now. Resonant tunneling through single ionized donor potentials is responsible for electron transport in disordered thin insulating barriers. This has been studied in the deeply insulating regime of large thin barriers formed by depleting electrostatically a semiconductor under a gate , or in thin amorphous silicon tunnel barriers . Interaction between distant impurity states in the channel have been revealed by peculiarities of the nonlinear transport. However, a single ionized donor potential cannot accomodate many electrons without becoming screened. We report for the first time Coulomb oscillations in very short MOSFETs with source drain distance $`d_{SD}`$ less than $`0.05\mu `$m and width much larger than $`d_{SD}`$. Contrarily to quasi 1D wires, where the size of the IQD is somewhat arbitrary, we will show that the diameter of the IQD is comparable to the source drain distance and that on resonance the conductance g in quantum units $`e^2/h`$ is close to one. Furthermore we demonstrate that the fluctuations of the conductance characterized by $`var(lng)`$ are proportional to the mean value $`<lng>`$, rather independent of geometry. Such an observation is consistent with one-parameter noninteracting scaling models of the metal-insulator transition. Our experimental findings suggest that interactions do not destroy this one-parameter description. The devices are MOSFETs on the (100) surface of silicon doped to a level of $`\mathrm{7\; 10}^{12}Borons/cm^2`$ for the 100nm series and of $`\mathrm{3\; 10}^{13}Borons/cm^2`$ for the 50nm series. The gate oxyde thickness is only $`d_{SiO_2}=3.8nm`$ for the 100nm series and $`d_{SiO_2}=2.4nm`$ for the 50nm series. This screens strongly Coulomb interactions in the inversion layer. Source and drain consist of highly ion implanted regions ($`10^{15}As/cm^2`$). The polysilicon gate has a length of 100nm or 50nm and the transverse dimension varies between 300 nm and 25 $`\mu m`$. The effective channel length $`d_{SD}`$ between source and drain is somewhat smaller than the geometrical gate length due to extension regions ($`\mathrm{2\; 10}^{14}As/cm^2`$). $`d_{SD}`$ is estimated to be of order 25 nm and 75 nm in the two cases considered here. The room temperature mobility is $`242cm^2s^1V^1`$ for the 100nm series and $`150cm^2s^1V^1`$ for the 50nm series. The source-drain current $`I_{SD}`$ is measured for a source-drain voltage $`V_{SD}=10\mu V`$ as a function of the gate voltage using a standard low-frequency lock-in technique. The $`I_{SD}V_{SD}`$ characteristics is linear for this value of $`V_{SD}`$ even at the lowest temperature, independent of $`V_g`$. The sample is inside a copper box thermally anchored to the mixing chamber of a dilution refrigerator. About 2 meters of Thermocoax Philips on each side of the sample provide the contact to the lock-in amplifier. Figure 1 shows the source-drain conductance in quantum units versus $`V_g`$ in three samples of gate length $`L=100nm`$ differing only by the width. At a temperature of $`T=35`$mK, we observe reproducible conductance fluctuations as a function of gate voltage which persist up to $`T=20`$K at small $`V_g`$. Depending on the value of the conductance these fluctuations evolve differently with temperature: If $`G>e^2/h`$ the fluctuations are gaussian around their mean value and poorly sensitive to temperature below T=4.2K. This characterizes universal conductance fluctuations in the diffusive regime not further considered here. If $`G<e^2/h`$, fluctuations evolve into sharp resonances at very low temperature (dielectric regime). In the diffusive regime, the linear increase of the conductance with the gate voltage permits to evaluate the mobility $`\mu `$: $`G_{square}=n_se\mu =C_gV_g\mu V_g`$ if $`\mu `$ and $`C_g`$ do not depend on $`V_g`$. We find $`\mu =25cm^2V^1s^1`$ for $`0.135VV_g0.2V`$ in the $`W=25\mu m`$ sample, $`\mu =53cm^2V^1s^1`$ for $`0.2VV_g0.3V`$ in the $`W=4\mu m`$ sample and $`\mu =73cm^2V^1s^1`$ for $`0.375VV_g0.6V`$ in the $`W=0.4\mu m`$ sample (at low temperature). These mobilities are weak reflecting a strong short range disorder and a mean free path of order $`4nm`$. The measured capacitance to gate coincides with the theoretical estimations: $`C_g/S=\frac{ϵ_0ϵ_{SiO_2}}{d_{SiO_2}}=10^{14}F/\mu m^2`$ (for $`d_{SiO_2}=3.8nm`$). A variation of $`\delta V_g=1V`$ induces $`\delta n_s=\mathrm{6.25\; 10}^{12}cm^2`$ for the 100nm series ($`\delta n_s=\mathrm{1.0\; 10}^{13}cm^2`$ for the 50nm series) . Typical conductance resonances in the dielectric regime are well fitted over 3 orders of magnitude (see inset of Fig.1) by the standard relation for the thermally broadened resonant tunneling regime: $`G(V_g,T)`$ $`=`$ $`{\displaystyle \frac{e^2}{h}}A{\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}L(V)\times {\displaystyle \frac{f(e\alpha (VV_g),T)}{V}}𝑑V`$ (1) $`L(V)`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Gamma }_e^2}{e^2\alpha ^2(V_0V)^2+\mathrm{\Gamma }_e^2}}`$ (2) with $`f(x,T)=(1+exp(x/k_BT))^1`$, $`\alpha =0.252`$ and $`A=0.8\pm 0.2`$. $`\mathrm{\Gamma }_e/k_B=60mK`$ is the intrinsic linewidth and $`V_0`$ is the gate voltage at the resonance. $`\delta E_F=e\alpha \delta V_g`$ is the variation of the Fermi energy. $`T=360mK\pm 15mK`$ is the effective electron temperature. $`A1`$ means that the localized state is equally connected to the source and the drain. If not, the resonant conductance would be exponentially smaller. From $`\alpha =C_g/e^2g_{2D}`$, we estimate an approximately constant density of states in the channel $`g_{2D}\mathrm{0.25\; 10}^{14}cm^2eV^1`$, which is reduced by a factor 4 compared to the metallic 2D density of states. This is due to Lifshitz tails induced by disorder at the bottom of the conduction band. The central result in Fig. 1 is that the onset of diffusive transport is essentially independent of the width of the sample. It always occurs when the total conductance $`Ge^2/h`$. This is also illustrated in a scaled representation of the fluctuations with respect to the average conductance $`<lng>`$ in the inset of Fig. 2 : Independent of geometry the log-normal fluctuations of lng disappear when $`<lng>0`$. The explanation of this striking result lies in the very broad distribution of conductances in the dielectric regime. Theories of log-normal distributions of conductances indicate that an anomalously large conductance, a rare event, dominates the total conductance. This is only valid at very low temperature, when parallel thermally activated conducting channels are exponentially small. According to the scaling theory of localization, at the metal-insulator transition, this event is the last conducting channel with conductance $`Ge^2/h`$ . In samples consisting of many squares in parallel and since the conductance distribution is log-normal, this last conducting channel carries almost the total current. Deep in the insulating regime, we find that the conductance fluctuations obey roughly to: $`var(lng)0.55<lng>`$, independent of geometry for widths up to $`W=4\mu m`$ (see Fig. 2). Here $`var(lng)=<[lng<lng>]^2>_{V_g}`$ is the variance of the average conductance $`<lng>`$ where the average is taken over $`V_g`$ . The proportionality between the variance and the mean of $`lng`$ is expected in disordered insulators at zero temperature and described by non-interacting scaling models. However, since interactions are crucial to explain the details of the fluctuations as shown below, our experiment suggests that a one-parameter description of the metal-insulator transition (at least in the strongly insulating phase) holds even in the presence of interactions. Fig. 3 shows Coulomb oscillations in our shortest samples with effective source drain distances $`d_{SD}`$ of 25nm and 75nm. In both samples each conductance peak at $`T=4.2K`$ corresponds to one peak at the lowest temperature. No additional peaks appear upon cooling and we only observe small shifts in gate voltage. This is consistent with single dot resonances and rules out multiple dot tunnelling as previously reported in quasi-1D wires. Actually, several dots acting in parallel explain qualitatively the complicated pattern observed in samples with widths larger than typically $`1\mu m`$. We first consider the series of resonances at relatively high gate voltage when the dot is filled with several electrons. Taking a simple Coulomb blockade model periodic conductance resonances occur if the temperature is lower than the charging energy $`E_Ck_BT`$. At zero source-drain voltage $`V_{SD}`$ and $`V_g`$ such that $`C_gV_g=Ne+e/2`$ where $`N`$ is the number of electrons in the dot, charge is transfered; If on the other hand $`V_{SD}=e/C`$, resonances occur at $`V_g`$ such that $`C_gV_g=Ne`$. This is demonstrated in the insets of Fig. 3. We suppose here that the capacitances to source and drain $`C_S=C_D=C/2`$ are equal. Within this classical model, we find $`C=\mathrm{8\; 10}^{17}F`$ and $`C_g=\mathrm{2\; 10}^{17}F`$ for the 100 nm sample, and $`C=\mathrm{4.6\; 10}^{17}F`$ and $`C_g=\mathrm{1\; 10}^{17}F`$ for the 50 nm sample. The charging energy is $`e^2/C2meV`$ and $`3.5meV`$ respectively for the 100nm and the 50nm sample. In the simplest model, $`\alpha =\delta E_F/e\delta V_g=C_g/(C_g+C)0.2`$, in agreement with the value deduced from the temperature dependence of the resonances (see inset of Fig. 1 ). With decreasing number of electrons on the dot, the mean level spacing $`\mathrm{\Delta }`$ increases and the barrier resistance between source and drain and the impurity quantum dot increases. If the temperature is below the mean level spacing we can estimate $`\mathrm{\Delta }`$ by the $`I_{SD}V_{SD}`$ non-linearities revealing the excited states of the dot. In our samples this is shown in Figure 4. Upon rising $`V_{SD}`$, the differential conductance $`\delta I_{SD}/\delta V_{SD}`$ exhibits a peak each time an excited state $`E_i`$ enters or exits the energy window $`\mu _S+\alpha eV_g>E_i>\alpha eV_g+\mu _D`$. Up to 3 excited states ( 6 peaks in $`\delta I_{SD}/\delta V_{SD}`$) are distinguishable for resonances at low gate voltage for $`V_{SD}3meV`$, such that $`\mathrm{\Delta }1meV`$. At higher gate voltage $`\mathrm{\Delta }`$ becomes smaller and the contrast between excited states is washed out (classical Coulomb blockade regime). Since $`\mathrm{\Delta }`$ is smaller but non negligible compared to $`E_C`$, the periodicity of the Coulomb oscillations is not strictly obeyed. The values for the mean spacing and charging energy are typical for all the measured samples. $`\mathrm{\Delta }1meV`$ corresponds to what is expected for single electron levels in a 2D box of $`65nm\times 65nm`$ taking into account the reduced density of states in the Lifshitz tail. Alternatively it corresponds also to a parabolic potential confinement of radius 50 nm, comparable to the source-drain distance, and heigth 10 meV, comparable to the Fermi energy at the MIT. The capacitance to gate $`C_g=\mathrm{2\; 10}^{17}F`$ (L=100nm) (resp. $`C_g=\mathrm{1\; 10}^{17}F`$ for L=50nm) corresponds to a 2D dot of size $`45nm\times 45nm`$ (resp. $`25nm\times 25nm`$). A main feature of our samples is the very thin gate oxyde which makes the gate very efficient to screen the Coulomb interaction in the IQD. For instance, a simple estimation of the repulsion between two bare electrons at a mean distance of $`r=20nm`$ in the channel gives $`V=\frac{1}{4\pi ϵ_0}2ϵ_{Si}\frac{d^2}{ϵ_{SiO_2}^2}\frac{e^2}{r^3}3.9meV`$ where $`ϵ_{Si}=11.4`$ and $`ϵ_{SiO_2}4`$ are the relative dielectric constants of $`Si`$ and $`SiO_2`$. Since the oxyde thickness ($`d_{SiO_2}=3.8nm`$) is much smaller than the distance between two electrons, the Coulomb interaction is a dipole interaction. The various above estimations confirm that the sequence of resonances reflects the interaction between electrons sharing the same (barely) localized site whose extension is comparable to the source-drain distance independent of the width of the channel or the dopand concentration. Furthermore, we have demonstrated in this way that in order to observe Coulomb blockade in nanostructures it is not necessary to constrict the current through a quasi 1D segment such as point contacts or wires. The domination of a rare event in a short 2DEG intrinsically favours a channel through a single IQD in not too wide geometries. In this way one has a better estimate of the relevant length and energyscales of the dot than in quasi 1D geometries. Also, downscaling the source-drain distance $`d_{SD}`$ from 75nm to 25nm increases the Coulomb energy, as expected when the dot size of the charge transmitting IQD decreases with channel length. In summary we have studied standard Si:MOSFETs of gate length $`L=50nm`$ and $`L=100nm`$ and various widths without any intentional confinement. Close to the MIT, characterized by $`Ge^2/h`$ at low temperature, conductance resonances as a function of the gate voltage are due to tunneling through a single disordered quantum dot, whose extension is comparable to the source-drain distance and which accomodates several electrons. In contrast to many previous studies on quasi-1d wires, our geometry allows to isolate single dot tunnelling in an impurity potential and quench multiple dot tunnelling. If the transverse dimension is increased too much, several IQD conduct for the same gate voltage ranges, causing eventually a complex structure of conductance resonances. Furthermore, we find that in the dielectric regime the variance of $`lng`$ systematically scales with the average of $`lng`$. Such an observation is consistent with a one-parameter description of the metal-insulator transition. Finally, our results strongly suggest that reducing the channel length even further should imply the observation of Coulomb oscillations with a charging energy comparable to room temperature.
no-problem/9908/hep-ph9908290.html
ar5iv
text
# QCD at Finite Density and Color Superconductivity ## 1 Brief history “Prehistoric” QCD-based calculations dealed with plasma-type phenomena like Debye screening etc , both for finite T and density. Early ideas about Color Superconductivity were based on simple observation: unlike electrons, quarks of different colors are attracted to each other even by Coulomb forces. Due to Cooper instability any small attraction is enough: however the superconducting gap was only $`\mathrm{\Delta }1MeV`$, and applicability of perturbative QCD was in doubt. My interest was initiated by finding that in the instanton liquid model even without $`any`$ quark matter, the ud scalar diquarks are very deeply bound, by amount comparable to the constituent quark mass. So, phenomenological manifestations of such diquarks have in fact deep dynamical roots: they follow from the same basic dynamics as the “superconductivity” of the QCD vacuum, the chiral ($`\chi `$-)symmetry breaking. These spin-isospin-zero diquarks are related to pions, and should be quite robust element of nucleon (octet baryons) structure<sup>1</sup><sup>1</sup>1As opposed to $`\mathrm{\Delta }`$ (decuplet) baryons. . Another argument for deeply bound diquarks comes from bi-color ($`N_c=2`$) theory: in it the scalar diquark is degenerate with pions. By continuity from $`N_c=2`$ to $`3`$, a trace of it should exist in real QCD<sup>2</sup><sup>2</sup>2 Instanton-induced interaction strength in diquark channel is $`1/(N_c1)`$ of that for $`\overline{q}\gamma _5q`$ one. It is the same at $`N_c=2`$, zero for large $`N_c`$, and is exactly in between for $`N_c=3`$.. Explicit calculations with instanton-induced forces for $`N_f=2,N_c=3`$ QCD have been made in two simultaneous<sup>3</sup><sup>3</sup>3Submitted to hep-ph on the same day. papers . Indeed, a very robust Cooper pairs and gaps $`\mathrm{\Delta }`$ 100 MeV were found. From then on, the field is booming. This phase (called CSC2) has the same symmetries as discussed before : the $`\chi `$-symmetry is restored but color group is broken by the diquark condensate, acting like Higgs VEV of the Standard Model. New variety of color superconductor, CSC3 with Color-flavor locking exists for 3 or more light flavors $`N_f=3`$, see sect.5. At asymptotically high densities the perturbation theory must become right, see section 6. Finally we will have discussion of some outstanding issues in sect.7. ## 2 Physics Overview The QCD phase diagram, as we understand it today, is shown in fig.1 at the $`\mu T`$ plane (baryonic chemical potential -temperature) At small $`T`$,$`\mu `$ there is ordinary hadronic matter with broken chiral symmetry. The point M (from “multi-fragmentation”) is the endpoint of the nuclear liquid-gas phase transition. At the (hypothetical) critical point E the first order line either continue as 2-nd order (for $`m_u,m_d=0`$) or disappears (for finite masses): according to recent proposal it can be found in real heavy ion collisions. QDQ (quark-diquark) phase is hypothetical : I have no time to speak about it here. The main point is locations of the two superconducting phases, CSC2 and CSC3. At T=0 going to large $`\mu `$, the $`\chi `$-symmetry seem to be first recovered in CSC2, and then broken again in CSC3. Let me then explain few major physics points. Why is there a transition from particle-hole to particle-particle pairing? The following figure (dispersion curves $`\omega (k)`$) for quarks in vacuum and superconductor explains it: it is better to have a gap at the surface of the Fermi sphere rather than the Dirac sea. Why instantons? The reasons are: (i) They are the strongest non-perturbative effect known; (ii) Unlike OGE, they $`do`$ explain quantitatively $`\chi `$-symmetry breaking in vacuum; (iii) Anomaly cannot be eliminated by finite density, so tunneling leads to level crossing at the surface of the Fermi sphere as well. Instantons create the following amusing triality: there are three attractive channels which compete: (i) the instanton-induced attraction in $`\overline{q}q`$ channel leading to $`\chi `$-symmetry breaking. (ii) the instanton-induced attraction in $`qq`$ which leads to color superconductivity. (iii) the light-quark-induced attraction of $`\overline{I}I`$, which leads to pairing of instantons into “molecules” and a Quark-Gluon Plasma (QGP) phase without $`any`$ condensates. How the calculations are actually done?. Analytically, mostly in the mean field approximation, similar to the original BCS theory in Gorkov formulation. Total thermodynamical potential consists of “kinetic energy” of the quark Fermi gas, including mass operators of two types (shown in figure below). The “potential energy” in such approximation is the interaction Lagrangian convoluted with all possible condensates. For example, instanton-induced one with $`N_f=3`$ leads to two types of diagrams shown in Fig.4, with (a) $`<\overline{q}q>^3`$ and (b) $`<qq>^2<\overline{q}q>`$. Then one minimizes the potential over all condensates and get gap equations: algebra may be involved because masses/condensates are color-flavor matrices. ## 3 Two colors: a very special theory One reason it is special is well known to lattice community: its fermionic determinant is $`real`$ even for non-zero $`\mu `$, which makes simulations possible. Early works by Karsch, Dagotto et al (of mid-80’s!) made sense, but were looked at only now. The major interest to this theory is related the so called Pauli-Gursey symmetry, due to which diquarks are degenerate with mesons. The $`\chi `$-symmetry breaking is $`SU(2N_f)Sp(2N_f)`$, for $`N_f=2`$ the coset $`K=SU(4)/Sp(4)=SO(6)/SO(5)=S^5`$. Those 5 massless modes are pions plus scalar diquark S and its anti-particle $`\overline{S}`$. The corresponding sigma model was worked out in : for further development see . As argued in , in this theory the critical value of transition to Color Superconductivity is simply $`\mu =m_\pi /2`$, the diquark condensate is just rotated $`<\overline{q}q>`$ one, and the gap is the constituent quark mass. Recent lattice works and instanton liquid simulation display it in great details, building confidence for other cases. ## 4 Two flavor QCD: the CSC2 phase This phase diagram is a rare example of calculated T-$`\mu `$ one: The 1st order line is dashed, and the 2nd order ones are solid lines. Most studies of this theory are at T=0. In all these works one more possible phase (intermediate between vacuum and CSC2), Fermi gas of constituent quarks, with both $`M,\mathrm{\Delta }0`$ \- was unstable. However in last more refined calculation it obtains a small window, as shown by the dashed line on the following figure. Its features are amusingly close to those of nuclear matter: but it isn’t, of course: to get nucleons one should go outside the mean field. First attempted to do so in was for another cluster - the $`\overline{I}I`$ molecules. At T=0 it is however only 10% correction to previous results, but is dominant as T grows. ## 5 $`N_f=3`$ QCD: the CSC3 phase The color-flavor locking means that diquark condensate has the structure $`q_i^aCq_j^b=\overline{\mathrm{\Delta }}_1\delta _{ia}\delta _{bj}+\overline{\mathrm{\Delta }}_2\delta _{ib}\delta _{ja}`$, where ij are color and ab flavor indices. It is very symmetric, reducing $`SU(3)_cSU(3)_fSU(3)_{diagonal}`$. It was verified in for the OGE interaction, and for instanton-induced one in : probably it is always true for that theory. Gaps $`\delta _i`$ and masses $`\sigma _i`$, following from instanton-based calculation , are shown as a function of $`\mu `$ in the following figure Two plus strange flavor QCD ($`m_s0`$) was studied in several papers . Just kinematically, us,ds Cooper pairs with zero momentum is difficult to make: for $`\mu _{u,d}=\mu _s`$ the momenta $`p_{u,d}^Fp_s^F`$. Instantons generate also dynamical operator $`m_s(\overline{u}\overline{d})(ud)`$. Resulting behavior is as shown in our first figure. ## 6 Asymptotically large densities At high densities $`\mu >1GeV`$ instantons are Debye-screened , as well as electric (Coulomb) OGE. So magnetic gluons overtake electric ones . $`Magnetically`$ bound Cooper pair is interesting by itself, as a rare example: one has to take care of time delay effects with Eliashberg eqn, etc. Angular integral leads to second log in the gap equation, leading to unusual answer: $`\mathrm{\Delta }\mu exp(3\pi ^2/\sqrt{2}g)`$ which implies that the gap grows indefinitely with $`\mu `$ <sup>4</sup><sup>4</sup>4 Numerical details for all densities can be found in recent work . and pQCD becomes finally justified. However, it is the case for huge densities, with $`\mu >10GeV`$ or so. ## 7 Physics issues under discussion The hadron-quark continuity. As pointed out in , the CSC3 phase not only has the same symmetries as hadronic matter (e.g. broken $`\chi `$-symmetry), but also very similar excitations. 8 gluons become 8 massive vector mesons, 3\*3 quarks become 8+1 “baryons”. The 8 massless pions remain massless<sup>5</sup><sup>5</sup>5Very exotic 3d objects, “super-qualitons” , the skyrmions made of pions are among the excitations.. Furthermore, photon and gluons are combined into a massless $`\gamma _{inside}`$. Can these phases be $`distiquished`$, and should there be $`any`$ phase transition (in $`N_f=3`$ theory)? Is it a superconductor, after all? I think the answers still is “yes”. For example, if one puts a piece of CSC3 into a magnet, it may levitate: although $`\gamma _{inside}`$ is massless, the magnet uses $`\gamma _{outside}`$ field and a part of it is expelled<sup>6</sup><sup>6</sup>6The same would happen with a small piece of Weinberg/Salam vacuum, if one can make magnet with “original” (“outside”) field.. Let me finish with few homework questions. What is the role of confinement in all these transitions? What is nuclear matter for different quark masses, anyway? Do we have other phases in between, like diquark-quark phase or (analog of) K condensation, or different crystal-like phases? Is there indeed a (remnant of) the tricritical point which we can find experimentally? And, above all, How to do finite density calculations on the lattice?
no-problem/9908/hep-ph9908302.html
ar5iv
text
# TRIVIALITY OF THE QUARK PROPAGATOR IN THE LADDER APPROXIMATION IN QCD ## I Introduction It is well known that the full dynamical information of any quantum field gauge theory such as QCD is contained in the corresponding quantum equations of motion, the so-called Schwinger-Dyson (SD) equations for propagators (lower Green’s functions) and vertices (higher Green’s functions) . The Bethe-Salpeter (BS) type integral equations for higher Green’s functions and bound-state amplitudes should be also included into this system. It should be complemented by the corresponding Slavnov-Taylor (ST) identities which, in general, relate lower and higher Green’s functions to each other. These identities are consequences of the exact gauge invariance and therefore $`\mathrm{"}areexactconstraintsonanysolutiontoQCD\mathrm{"}`$ . Precisely this system of equations can serve as an adequate and effective tool for the non-perturbative approach to QCD . It is important to understand however, that the above mentioned system is an infinite chain of strongly coupled highly nonlinear integral equations, so there is no hope for an exact solutions. For this reason, some truncation scheme is always needed in order to make these equations tractable for getting physical information from them. Because of its relative simplicity the most popular is the ladder approximation (LA) in various forms. In general, it consists of approximating the full vertices by their free perturbative (point-like) counterparts in the corresponding kernels of the above mentioned integral equations. Approximating in addition the full gluon propagator by its free perturbative expression, for example in the BS-type integral equation for the vertex, one gets the so-called quenched LA (QLA) scheme which for the quark SD equation became known as ”rainbow” approximation. However, very soon it was realized that QLA was too crude for QCD and should be improved in order to incorporate the QCD renormalization group results for the running coupling constant (asymptotic freedom) into the system of SD equations. The so-called improved LA (ILA) was proposed (special dependence of the full gluon form factor on its momentum). It makes it possible to take into account self-interacting gluon modes (non-Abelian character of QCD) at the level of the full gluon propagator only, so vertices remain intact, i. e. point-like ones. The main problem of the LA is, of course, its self-consistency. For a long time, it was widely believed that the quark sector could be effectively decoupled from other sectors in QCD in the LA. For example, it is usually assumed that ghost degrees of freedom are not important for the above mentioned SD system of equations in this case. At first sight this is so indeed, since neither the quark SD equation nor the equation for the quark-gluon vertex depends explicitly on ghosts in the LA. However, in QCD as it was mentioned above, these two equations should be complemented by the corresponding ST identity which explicitly depends on ghost degrees of freedom (even in the LA). Unlike QED, QCD is much more complicated gauge theory. It contains many different sectors. Making some truncation in one sector, it is necessary to be sure that nothing is going wrong in other sectors since the above mentioned SD system of equations may remain strongly coupled even after truncation. The key elements relating different sectors in QCD are the above mentioned ST identities, playing thus an important role in the investigation of the problem of self-consistency of any truncation scheme. In other words, any solution to QCD, in particular LA should be compatible with these identities and not vice versa as it has been emphasized above. Especially important the problem of self-consistency of the LA becomes when it is applied to such essentially non-perturbative problems as quark confinement and dynamical chiral symmetry breaking (DCSB). Its inconsistency for the above mentioned problems was pointed out by Adler who explicitly showed (within precisely the dynamical equations approach) that a more sophisticated approximation than the LA is needed to understand quark confinement and DCSB in QCD. The importance of the self-consistent treatment of DCSB within the SD system of equations in the LA was, apparently for the first time, emphasized in Ref. . The next interesting step in this direction was done quite recently in Ref. where a systematic method was developed for obtaining consistent approximations to the SD and BS equations which maintained the external gauge invariance (by using the Ward-Takahashi (WT) identities for the color-singlet, but flavor-nonsiglet vertices). In contrast to the above mentioned paper , the main purpose of this paper is to investigate the self-consistency of the LA by using the ST identity for the color-nonsiglet, but flavor-singlet quark-gluon vertex in QCD (the internal gauge invariance). Precisely this vertex is one of the most important key objects in QCD itself. Everybody knows that the LA is bad (for example, it is explicitly gauge-dependent truncation scheme), nevertheless eveybody continues to use it. In some sense it was justified (see, for example Ref. ) since up until now there was not an exact criterion to prove or disprove the LA in gauge theories. Here we precisely propose how to formulate this criterion. In section 2 we remind some well-known things from QED in order to better explain our method. In section 3 we derive the constraint equation and in sections 4 and 5 it is solved without and with ghosts, respectively requiring the full qurk propagator to be almost free one. In section 6 the flavor-nonsinglet but color-singlet axial-vector vertex is investigated. Some general remarks on the non-perturbative renormalization in the Landau gauge is presented in section 7. Our conclusions are given in section 8. ## II QED Let us start with some well-known things from QED and consider first unrenormalized (for simplicity) SD equation in the LA for the quark (electron) propagator in momentum space $$S^1(p)=S_0^1(p)g^2\frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\gamma _\beta D_{\alpha \beta }(q),$$ (1) where $`S_0^1(p)=i(\widehat{p}m_0)`$ with $`m_0`$ being the current (”bare”) mass of a single quark (electron) and $`D_{\alpha \beta }(q)`$ is the full gluon (photon) propagator in the arbitrary covariant gauges, $$D_{\alpha \beta }(q)=i\left\{\left[g_{\alpha \beta }\frac{q_\alpha q_\beta }{q^2}\right]\frac{1}{q^2}d(q^2;\xi )+\xi \frac{q_\alpha q_\beta }{q^4}\right\}$$ (2) and $`\xi `$ is the gauge-fixing parameter ($`\xi =0`$, Landau gauge). Here $`q=pl`$ is the transfer momentum and we assign the factor $`ig\gamma `$ to the point-like (”bare”) veritices with the corresponding Dirac indices. Differentiating Eq. (2.1) with respect to $`p_\mu `$, one obtains $`_\mu S^1(p)`$ $`=`$ $`i\gamma _\mu _\mu g^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\gamma _\beta D_{\alpha \beta }(q)}`$ (3) $`=`$ $`i\gamma _\mu +g^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu S(l)]\gamma _\beta D_{\alpha \beta }(q)},`$ (4) up to unimportant total derivative (which is assumed as usual to vanish at the ends of integration). This is the differential form of the quark SD equation (2.1) relevant for further discussion in the next section as well. Let us consider now the Bethe-Salpeter (BS) integral equation in the LA for the quark-photon vertex $`\mathrm{\Gamma }_\mu (p,q)`$ ($`q=p^{}p`$) at zero momentum transfer $`q=0`$, $$\mathrm{\Gamma }_\mu (p,0)=i\gamma _\mu g^2\frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\mathrm{\Gamma }_\mu (l,0)S(l)\gamma _\beta D_{\alpha \beta }(q).$$ (5) The corresponding WT identity in QED, $`q_\mu \mathrm{\Gamma }_\mu (p+q,p)=S^1(p+q)S^1(p)`$, provides an exact solution for $`\mathrm{\Gamma }_\mu (p,0)`$, namely $$\mathrm{\Gamma }_\mu (p,0)=_\mu S^1(p).$$ (6) Substituting it into the Eq.(2.4) and using the obvious identity $`_\mu S^1(p)=S^1(p)[_\mu S(p)]S^1(p)`$, one finally obtains $$_\mu S^1(p)=i\gamma _\mu +g^2\frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu S(l)]\gamma _\beta D_{\alpha \beta }(q).$$ (7) This is the quark SD equation obtained from the BS integral equation for the corresponding vertex on account of the WT identity (2.5). Comparing it with that of Eq. (2.3) (second line), one immediately concludes that they are identical. The renormalized version of Eq. (2.3) is obtained by the multiplication of its inhomogenious term by the quark wave function renormalization constant $`Z_2`$, while the renormalized version of Eq. (2.6) is obtained by the multiplication of its inhomogenious term by the vertex renormalization constant $`Z_1`$. Because of the WT identity (2.5) these constants are equal to each other ($`Z_1=Z_2`$) and again renormalized versions of these equations coincide in QED. So the BS equation at zero momentum transfer and on account of the corresponding WT identity (which is a consequence of the exact gauge invariance) does not provide an additional constraint on the quark SD equation in QED. In other words, the summation of the ladder diagrams within the corresponding BS integral equation at zero momentum transfer on account of the WT identity is consistent with the electron (quark) SD equation itself. ## III QCD The situation in QCD is completely different and is much more complicated because of color and ghost degrees of freedom which precisely provide an addition constraint on the solution to the quark SD equation in the LA by using the same method as in previous section for QED. It is convenient to begin our analysis from the quark SD equation in QCD which is just the same as in QED (2.1) with only trivial replacement $`g^2g_F^2`$ because of color group factors. Here $`g_F^2=g^2C_F`$ and $`C_F`$ is the eigenvalue of the quadratic Casimir operator in the fundamental representation (for SU(N), in general, $`C_F=(N^21)/2N=4/3,N=3`$). Thus the renormalized version of the quark SD equation (2.1) is as follows $`\stackrel{~}{S}^1(p)`$ $`=`$ $`Z_2S_0^1(p)+i\stackrel{~}{\mathrm{\Sigma }}(p)`$ (8) $`=`$ $`Z_2S_0^1(p)\stackrel{~}{g}_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha \stackrel{~}{S}(l)\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q)},`$ (9) where the renormalized quark propagator and the full gluon propagator (2.2) are related to the unrenormalized ones as $`S(p)`$ $`=`$ $`Z_2\stackrel{~}{S}(p),`$ (10) $`D_{\alpha \beta }(q)`$ $`=`$ $`Z_V\stackrel{~}{D}_{\alpha \beta }(q)`$ (11) with the renormalization of the coupling constant is beeing determined by the corresponding combination of the above and below introduced renormalization constants, $`Z^{}s`$ . The differential form of the renormalized quark SD equation becomes (compare with Eq. (2.3)) $`_\mu \stackrel{~}{S}^1(p)`$ $`=`$ $`Z_2i\gamma _\mu +_\mu i\stackrel{~}{\mathrm{\Sigma }}(p)`$ (12) $`=`$ $`Z_2i\gamma _\mu +\stackrel{~}{g}_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu \stackrel{~}{S}(l)]\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q)}.`$ (13) Let us introduce now the renormalized ST identity $$k_\mu \stackrel{~}{\mathrm{\Gamma }}_\mu ^a(p,k)\left[Z_g^1+\stackrel{~}{b}(k^2;\xi )\right]=\left[Z_B^1T^a\stackrel{~}{B}^a(p,k;\xi )\right]\stackrel{~}{S}^1(p+k)\stackrel{~}{S}^1(p)\left[Z_B^1T^a\stackrel{~}{B}^a(p,k;\xi )\right],$$ (14) where the corresponding renormalized quark-gluon vertex, the ghost self-energy and the ghost-quark scattering kernel<sup>*</sup><sup>*</sup>*For its skeleton expansion see, for example papers and references therein. In the LA only the first term of this expansion survives with the replacement of the corresponding full ghost- and quark-gluon vertices by their free perturbative (point-like) counterparts. However, there is no need to exploit the LA explicit expressions for quantities defined in (3.8) and (3.9) (see below) within our approach. are related to their unrenormalized counterparts as follows: $`\mathrm{\Gamma }_\mu (p,k)`$ $`=`$ $`Z_1^1\stackrel{~}{\mathrm{\Gamma }}_\mu (p,k),`$ (15) $`b(k^2;\xi )`$ $`=`$ $`Z_g\stackrel{~}{b}(k^2;\xi ),`$ (16) $`B(p,k;\xi )`$ $`=`$ $`Z_B\stackrel{~}{B}(p,k;\xi ).`$ (17) For furture aim we have introduced an explicit dependence on a gauge fixing parameter $`\xi `$ into the ghost degrees of freedom. Let us note that in the ST identity (3.3) the transfer momentum runs through the ghost degrees of freedom. The following important relation between renormalization constants holds $$Z_1Z_B=Z_2Z_g.$$ (18) It is worthwhile noting that for the sake of furture convenience instead of the ghost field renormalization constant and the ”renormalization” constant $`Y`$ for the amplitude $`H(p,k)=1B(p,k)`$, defined in Ref. , we have introduced the ghost self-energy remormalization constant $`Z_g`$ and the ”renormalization” constant $`Z_B`$ for the ghost-quark scattering kernel itself. Differentiation of the renormalized ST identity (3.3) with respect to $`k_\mu `$ and then setting $`k=0`$, yieldsIt is assumed that possible unphysical kinematic singularities have been already removed from the vertex by means of Ball and Chiu procedure . $$\stackrel{~}{\mathrm{\Gamma }}_\mu ^a(p,0)\left[Z_g^1+\stackrel{~}{b}(0;\xi )\right]=Z_B^1T^a_\mu \stackrel{~}{S}^1(p)\stackrel{~}{\mathrm{\Pi }}_\mu ^a(p,0;\xi )\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )\stackrel{~}{S}^1(p)+\stackrel{~}{S}^1(p)\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi ),$$ (19) where we introduced the following notations $$\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )=\left[\frac{}{k_\mu }\stackrel{~}{B}^a(p,k;\xi )\right]_{k=0}$$ (20) and $$\stackrel{~}{\mathrm{\Pi }}_\mu ^a(p,0;\xi )=\stackrel{~}{B}^a(p,0;\xi )_\mu \stackrel{~}{S}^1(p).$$ (21) Here a few remarks are in order. Though the regular dependence of the ghost-quark scattering kernel $`\stackrel{~}{B}^a(p,k;\xi )`$ on the ghost self-energy momentum $`k`$ is preserved by the general Taylor’s result $$\stackrel{~}{B}^a(p,0;\xi )=\xi F^a(p)=0at\xi =0,$$ (22) which shows that this kernel exists at small $`k`$ in any covariant gauge and vanishes only in the Landau gauge $`\xi =0`$, nevertheless the regular dependence of the ghost self-energy itself on its momentum is not so obvious. However, here we treat the ghost-self-energy as a regular function of its momentum since a singular dependence (which, in principle should not be excluded $`apriori`$) requires completely different investigation and is left for consideration elsewhere. The constraint equation in this case also will be completely different from that obtained below. The renormalized version of the BS-type integral equation for the color non-singlet vertex at zero momentum transfer is $$\stackrel{~}{\mathrm{\Gamma }}_\mu ^a(p,0)=Z_1i\gamma _\mu T^a\stackrel{~}{g}^2\frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Gamma }}_\mu ^a(l,0)\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q).$$ (23) Substituting (3.7) into the BS-type integral equation (3.11), one obtains $`Z_B^1T^a_\mu \stackrel{~}{S}^1(p)\stackrel{~}{\mathrm{\Pi }}_\mu ^a(p,0;\xi )`$ $``$ $`\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )\stackrel{~}{S}^1(p)+\stackrel{~}{S}^1(p)\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )=Z_1[Z_g^1+\stackrel{~}{b}(0;\xi )]i\gamma _\mu T^a`$ (24) $`+`$ $`\stackrel{~}{g}^2Z_B^1T^bT^aT^b{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu \stackrel{~}{S}(l)]\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q)}`$ (25) $`+`$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Pi }}_\mu ^a(l,0;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}`$ (26) $``$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{\mathrm{\Psi }}_\mu ^a(l;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}`$ (27) $`+`$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Psi }}_\mu ^a(l;\xi )\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)},`$ (28) where the obvious identity, $`_\mu S^1(p)=S^1(p)[_\mu S(p)]S^1(p)`$, has been already used. Let us now use the commutation relation between color matricies $`[T^a,T^b]=if_{abc}T^c`$, where $`f_{abc}`$ are the antisymmetric $`SU(3)`$ structure constants with non-zero values, given for example in Ref. . Then one obtains, $$T^bT^aT^b=[C_F\frac{1}{2}C_A]T^a,$$ (29) where $`C_F`$ is the above mentioned eigenvalue of the quadratic Casimir operator in the fundamental representation while $`C_A`$ is the same but in the adjoint representation, $`C_A=N`$ for $`SU(N)`$ ($`N=3`$ for QCD). So from this relation and on account of the renormalized version of the differential form of the quark SD equation (3.3) (because of $`\stackrel{~}{g}^2C_F=\stackrel{~}{g}_F^2`$) and using finally an important relation (3.6), one arrives at $`\stackrel{~}{\mathrm{\Pi }}_\mu ^a(p,0;\xi )`$ $``$ $`\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )\stackrel{~}{S}^1(p)+\stackrel{~}{S}^1(p)\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )=Z_1\stackrel{~}{b}(0;\xi )i\gamma _\mu T^a`$ (30) $``$ $`{\displaystyle \frac{1}{2}}C_AT^a\stackrel{~}{g}^2Z_B^1{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu \stackrel{~}{S}(l)]\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q)}`$ (31) $`+`$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Pi }}_\mu ^a(l,0;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}`$ (32) $``$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{\mathrm{\Psi }}_\mu ^a(l;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}`$ (33) $`+`$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Psi }}_\mu ^a(l;\xi )\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}.`$ (34) This is a renormalized version of the general constraint equation which relates ghost degrees of freedom to those of quark ones in a new manner within the LA. In principle, the solution (if any) of the quark SD equation (3.1) itself should be compatible with this constarint equation. At first sight it is too complicated integral equation, but nevertheless its analysis is rather simple. Here let us note that constraint (3.14) is slightly simplified in the Landau gauge $`\xi =0`$ since the composition (3.9) vanishes due to the above mentioned Taylor’s general result (3.10). ## IV Solution without ghosts As it has been already mentioned in the Introduction, it is widely believed that the ghost degrees of freedom are not important for the LA to covariant gauge QCD. If this is so indeed, let us omit them from the general constraint equation (3.14) by ”hand”, which is equivalent to consider QED-type ST identity (3.4) (apart from the color group generators) from the very beginning. Apart from neglecting corrections to the vertex, this is the standard procedure in the LA to covariant gauge QCD. It has been already done in order to proceed from the SD equation for the vertex to its BS-type counterpart (3.11) which allows precisely to explicitly sum up ladder contributions (see, for example Ref. ). Doing so in (3.14) and on account of Eq. (3.3), one obtains $$\frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu \stackrel{~}{S}(l)]\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q)=\stackrel{~}{g}_F^2_\mu i\stackrel{~}{\mathrm{\Sigma }}(p)=0,$$ (35) which implies general solution is to be $`i\stackrel{~}{\mathrm{\Sigma }}(p)=im_c`$, where $`m_c`$ is the constant of integration of the dimension of mass. From the quark SD equation (3.1) (first line) $$\stackrel{~}{S}^1(p)=Z_2S_0^1(p)+i\stackrel{~}{\mathrm{\Sigma }}(p)$$ (36) with $`S_0^1(p)=i(\widehat{p}m_0)`$ and on account of the above mentioned trivial solution to the quark self-energy, it finally follows $$\stackrel{~}{S}(p)=\frac{i\stackrel{~}{Z}_2}{\widehat{p}\stackrel{~}{m}},$$ (37) where $`\stackrel{~}{Z}_2^1=Z_2`$ and $`\stackrel{~}{m}=m_0+\stackrel{~}{Z}_2m_c`$. Thus the solution of the constraint equation (3.14) with ghosts omitted by ”hand” in the LA to covariant gauge QCD requires that the quark propagator should be almost free one apart from the redefinition of the quark mass. In other words, there should be $`nonontrivial`$ solution to the quark SD equation in the LA in this case. So the main question arises, namely may ghosts cure this problem or not? The final answer is, of course, not (see next section). If they would cure this problem it would mean that the quark propagator (via the general constraint equation (3.14)) would explicitly depend on ghost degrees of freedom. However, this is impossible in QCD where ghosts contribute only into the closed loops, so nothing can explicitly depend on them exept of the contributions to the renormalization constants, constants of integration, etc. Precisely the above mentioned problem will be investigated below in more sophisticated fashion. ## V Solution with ghosts In order to investigate the general constraint equation (3.14) in more sophisticated way, let us write down formal BS-type integral equations for the compositions defined in (3.8) and (3.9). Like the quark-gluon vertex, these objects are vector quantities at zero momentum transfer. They satisfy formally absolutely the same BS-type integral equations in the LA as the above mentioned quark-gluon vertex at zero momentum transfer (3.11). Thus they are the LA equations for (3.8) and (3.9) in the same way as (3.11) is the LA to the exact BS integral equation for the quark-gluon vertex at zero momentum transfer. The one of the differences is that there are no point-like counterparts of these quantites in QCD. But the main difference is that, in contrast to the BS-type integral equation for the quark-gluon vertex (3.11), these equations are completely auxiliarly. They have no independent role, the main purpose to use (or to postulate) them is to show (as was mentioned above) that nothing explicitly depends on ghost degrees of freedom in QCD. So in complete analogy with (3.11), the formal homogenious BS-type equation in the LA for the composition (3.9) looks like $$\stackrel{~}{\mathrm{\Pi }}_\mu ^a(p,0;\xi )=\stackrel{~}{g}^2\frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Pi }}_\mu ^a(l,0;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q).$$ (38) For the composition defined in (3.8) it is convenient to write down the formal BS-type integral equations for its left and right hand side combinations, $`\stackrel{~}{\mathrm{\Psi }}_\mu ^{a(l)}(p;\xi )=\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )\stackrel{~}{S}^1(p)`$ and $`\stackrel{~}{\mathrm{\Psi }}_\mu ^{a(r)}(p;\xi )=\stackrel{~}{S}^1(p)\stackrel{~}{\mathrm{\Psi }}_\mu ^a(p;\xi )`$, respectively $`\stackrel{~}{\mathrm{\Psi }}_\mu ^{a(l)}(p;\xi )`$ $`=`$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Psi }}_\mu ^{a(l)}(l;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}`$ (39) $`\stackrel{~}{\mathrm{\Psi }}_\mu ^{a(r)}(p;\xi )`$ $`=`$ $`\stackrel{~}{g}^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha T^b\stackrel{~}{S}(l)\stackrel{~}{\mathrm{\Psi }}_\mu ^{a(r)}(l;\xi )\stackrel{~}{S}(l)\gamma _\beta T^b\stackrel{~}{D}_{\alpha \beta }(q)}.`$ (40) Substituting (5.1) and (5.2) into the constraint equation (3.14), we are left with $$Z_1\stackrel{~}{b}(0;\xi )i\gamma _\mu =\frac{1}{2}C_A\stackrel{~}{g}^2Z_B^1\frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu \stackrel{~}{S}(l)]\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q).$$ (41) From Eq. (3.3), however it follows $$\frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu \stackrel{~}{S}(l)]\gamma _\beta \stackrel{~}{D}_{\alpha \beta }(q)=\stackrel{~}{g}_F^2_\mu i\stackrel{~}{\mathrm{\Sigma }}(p),$$ (42) so Eq. (5.3) by denoting $`\stackrel{~}{b}_1(0;\xi )=2Z_1Z_B(C_F/C_A)\stackrel{~}{b}(0;\xi )=2Z_2Z_g(C_F/C_A)\stackrel{~}{b}(0;\xi )`$, because of (3.6), becomes $`i\gamma _\mu \stackrel{~}{b}_1(0;\xi )=_\mu i\stackrel{~}{\mathrm{\Sigma }}(p)`$. Its general solution is, $`i\stackrel{~}{\mathrm{\Sigma }}(p)=i\widehat{p}\stackrel{~}{b}_1(0;\xi )+im_c`$, where $`m_c`$ is the constant of integration of the dimension of mass. From the quark SD equation (4.2) it finally follows $$\stackrel{~}{S}(p)=\frac{i\stackrel{~}{Z}_2}{\widehat{p}\stackrel{~}{m}},$$ (43) where $`\stackrel{~}{Z}_2^1`$ $`=`$ $`Z_2\left(1+2Z_g(C_F/C_A)\stackrel{~}{b}(0;\xi )\right),`$ (44) $`\stackrel{~}{m}`$ $`=`$ $`\stackrel{~}{Z}_2(m_c+Z_2m_0).`$ (45) Thus the use of the auxiliarly BS-type integral equations (5.1) and (5.2) allows one to treat ghost degrees of freedom in more sophisticated way than simply omit them by ”hand” (as in previous section), but, nevertheless the quark propagator in the LA again remains trivial one (5.5) apart from the redefinitions of the quark mass and the quark wave function renormalization constant with the help of ghost self-energy at zero point (compare with (4.3)). Nothing explicitly depends on ghosts as it should be indeed in QCD. It is well known that in the Landau gauge $`\xi =0`$ the quark wave function is not renormalized, i. e. $`Z_2=1`$. Whether the redefined renormalization constant $`\stackrel{~}{Z}_2`$ is also will be equal to unity in this gauge depends on the solution of the SD equation for the ghost self-energy. This investigation as well as the solution of the quark SD equation itself in order to determine a possible value of the constant of integration $`m_c`$ is, obviously, beyond the scope of this paper and is left for consideration elsewhere (see also section 7). ## VI QCD. Flavor-nonsinglet axial-vector vertex It is convenient to continue our discussion of the situation in QCD with the ST identity for the flavor-nonsinglet (but color-singlet), axial-vector vertex. It is not complicated by the unknown ghost contributions. In QED there is no sense to consider axial-vector vertex because there exists only flavor-singlet channel which, obviously, will be always complicated by the axial anomaly . Contrast to QED, in QCD there exists flavor-nonsinglet channel which is free from axial anomaly, but a new interesting problem occurs with this identity. Also not loosing generality, let us consider the chiral limit ($`m_0=0`$) of this identity since the nonchiral case, evidently, can not change conclusions drawn below. Let us consider the flavour non-singlet, axial-vector WT identity in the chiral limit $$iq_\mu \mathrm{\Gamma }_{5\mu }^i(p+q,p)=\left(\frac{\lambda ^i}{2}\right)\{\gamma _5S^1(p)+S^1(p+q)\gamma _5\},$$ (46) where $`q=p^{}p`$ is the momentum transfer (the external momentum) and the quark propagator is given by $$iS(p)=\widehat{p}A(p^2)+B(p^2),$$ (47) so its inverse is expressed as $$\{iS(p)\}^1=\widehat{p}\overline{A}(p^2)\overline{B}(p^2),$$ (48) with $`\overline{A}(p^2)`$ $`=`$ $`A(p^2)D^1(p^2),`$ (49) $`\overline{B}(p^2)`$ $`=`$ $`B(p^2)D^1(p^2),`$ (50) $`D(p^2)`$ $`=`$ $`p^2A^2(p^2)B^2(p^2).`$ (51) In connection with (6.1), one has to point out that, in general, $`\lambda ^i`$ is a $`SU(N_f)`$ flavour matrix and, in the massless case, the quark propagator is proportional to the unit matrix in the flavour space. It is well known that if dynamical chiral symmetry breaking (DCSB) at the quark level, implemented as $$\{S^1(p),\gamma _5\}_+=i\gamma _52\overline{B}(p^2)0$$ (52) takes place, then the axial-vector vertex should have a pole corresponding to a Goldstone state. Indeed, from (6.5) it follows that $`\mathrm{\Gamma }_{5\mu }^i(p+q,p)`$ has a pseudoscalar pole at $`q^2=0`$ (dynamical singularity which determines Goldstone state) if and only if the double of the dynamically generated quark mass function $`2\overline{B}(p^2)`$ in (6.5) is nonzero and vice versa. In order to self-consistently untangle regular and pole parts in (6.1), it is necessary to perform a procedure described in detail in our paper . It is a similar to that of the above mentioned Ball and Chiu procedure to remove kinematic (unphysical) singularities from the vertex. As a result of this, dependence on the two arbitrary form factors occurs in the regular part of this vertex at zero momentum transfer $`q=0`$, namely $$\mathrm{\Gamma }_{5\mu }^{iR}(p,p)=\left(\frac{\lambda ^i}{2}\right)[i_\mu S^1(p)+\mathrm{\Delta }_\mu (p)]\gamma _5,$$ (53) where $$\mathrm{\Delta }_\mu (p)=\gamma _\mu R_6(p^2)\widehat{p}\gamma _\mu R_{11}(p^2).$$ (54) The corresponding BS integral equation becomes equation for these form factors and therefore no statement about justification of the LA can be deduced. Indeed, let us show this explicitly. The BS integral equation for the regular part of the axial-vector vertex at zero momentum transfer is (here and below in this section again $`q=pl`$) $$\mathrm{\Gamma }_{5\mu }^{iR}(p,p)=\left(\frac{\lambda ^i}{2}\right)\gamma _5\gamma _\mu g_F^2\frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\mathrm{\Gamma }_{5\mu }^{iR}(l,l)S(l)\gamma _\beta D_{\alpha \beta }(q),$$ (55) where summation over color indices is already done, i. e. $`g_F^2=g^2C_F`$ (see Eq. (3.1)). Substituting relations (6.6-6.7) into this equation and on account of the above used identity, one obtains $`[_\mu S^1(p)+i\mathrm{\Delta }_\mu (p)]\gamma _5`$ $`=`$ $`i\gamma _5\gamma _\mu `$ (56) $`+`$ $`g_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu S(l)]S^1(l)\gamma _5S(l)\gamma _\beta D_{\alpha \beta }(q)}`$ (57) $``$ $`ig_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\mathrm{\Delta }_\mu (l)\gamma _5S(l)\gamma _\beta D_{\alpha \beta }(q)}.`$ (58) Let us now introduce the following notation $$S^1(p)\gamma _5S(p)=\gamma _5+\mathrm{\Delta }(p)\gamma _5.$$ (59) where, obviously $`\mathrm{\Delta }(p)`$ can be given in terms of the quark propagator form factors (6.4). Multiplication of Eq. (6.9) by $`\gamma _5`$ from right and on account of (6.10), yields $`_\mu S^1(p)+i\mathrm{\Delta }_\mu (p)`$ $`=`$ $`i\gamma _\mu `$ (60) $`+`$ $`g_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu S(l)]\gamma _\beta D_{\alpha \beta }(q)}`$ (61) $``$ $`g_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu S(l)]\mathrm{\Delta }(l)\gamma _\beta D_{\alpha \beta }(q)}`$ (62) $``$ $`ig_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\mathrm{\Delta }_\mu (l)\gamma _5S(l)\gamma _\beta \gamma _5D_{\alpha \beta }(q)}.`$ (63) Using futher the QCD quark SD equation (3.3), one finally obtains $`i\mathrm{\Delta }_\mu (p)=`$ $``$ $`g_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha [_\mu S(l)]\mathrm{\Delta }(l)\gamma _\beta D_{\alpha \beta }(q)}`$ (64) $``$ $`g_F^2{\displaystyle \frac{d^nl}{(2\pi )^n}\gamma _\alpha S(l)\mathrm{\Delta }_\mu (l)\gamma _5S(l)\gamma _\beta \gamma _5D_{\alpha \beta }(q)},`$ (65) where, recalling $`\mathrm{\Delta }_\mu (p)`$ is given in (6.7). This is the inhomogenious BS integral equation for unknown form factors $`R_6`$ and $`R_{11}`$, so, as was mentioned above, no statement about justification of the LA can be deduced. Evidently, this is a result of the fact that the exact expression for the longitudinal part of color-singlet but flavor-nonsinglet axial-vector ST identity is known up to two arbitrary form factor even at zero momentum transfer. Also obviously, that one can derive the similar BS equation for these form factors in some other approximation for the BS scattering kernel and again no statement can be achived in favour of this or another truncation scheme. ## VII Remarks on the non-perturbative renormalization in the Landau gauge In this section we are not restricted to the LA and are going to make a few general remarks concerning the non-perturbative renormalization in the Landau gauge. It is well-known that in QCD the renormalization program at zero point is greatly simplified in the Landau gauge. The ghost degrees of freedom, such as the ghost-gluon vertex and the ghost-quark scattering kernel, do not renormalize at zero point in the Landau gauge. Their renormalization constants (defined at zero point) become trivial unites in this gauge. Summary of these results one can find in Ref. . This fails if the renormalization is performed at nonzero euclidean mass scale parameter (($`\mu ^2`$), renormalization point) because of the possible infrared (IR) singularities due to exchange of massless gluons. However, as was mentioned above, the IR singular behaviour of the ghost self-energy requires special treatment and will be considered elsewhere.The treatment of ghost degrees of freedom in the Landau gauge in Ref. is not completely self-consistent. It was not investigated whether the enhancement of the ghost propagator in the IR was compatible with Taylor’s general result (3.10) for the ghost-quark scattering kernel which was simply omitted by hand in their truncation scheme. Let us consider now the exact ghost propagator and its self-energy which are related to each other as follows $$G(k)=\frac{i}{k^2[1+b(k^2;\xi )]},$$ (66) where, for simplicity’s sake, we suppressed the color group indicies. Introducing the renormalized ghost propagator $$G(k)=Z_G\stackrel{~}{G}(k),$$ (67) from (7.1-7.2) and (3.5), one obtains $$Z_G=Z_g^1$$ (68) and $$\stackrel{~}{G}(k)=\frac{i}{k^2[1+b^R(k^2;\xi )]},$$ (69) where we define as usual (see, for example, Ref. ) $$b^R(k^2;\xi )=\stackrel{~}{b}(k^2;\xi )\stackrel{~}{b}(0;\xi )$$ (70) and $$Z_g^1+\stackrel{~}{b}(0;\xi )=1.$$ (71) In a similar way let us define the regularized quark-ghost scattering amplitude (3.5) as follows $$B^R(p,k;\xi )=\stackrel{~}{B}(p,k;\xi )\stackrel{~}{B}(0,0;\xi )$$ (72) and $$Z_B^1\stackrel{~}{B}(0,0;\xi )=1,$$ (73) so that $$Z_B^1\stackrel{~}{B}(p,k;\xi )=1B^R(p,k;\xi )$$ (74) From Taylor’s general result (3.10) and on account of (7.8) then it follows that in the Landau gauge $`\xi =0`$, one has $$Z_B=1,$$ (75) and consequently from (3.6) it follows $$Z_1=Z_2Z_g.$$ (76) It is worthwhile noting that the renormalization constant $`Z_B`$ coincides with the renormalization constant $`Y`$ defined in Ref. for the amplitude $`H(p,q)=1B(p,q)`$, i.e. $`Z_BY`$ in the Landau gauge $`\xi =0`$ at least. Let us now briefly discuss a possible interesting feature of the renormalization at zero point in the Landau gauge. Indeed, from (7.11) it follows that if the ghost self-energy renormalization constant $`Z_g`$ (defined at zero point) would may equal to unity as $`Z_B`$ (7.10) then the QED-type relation between the QCD renormalization constants (defined at zero in the Landau gauge) will be recovered, namely (7.11) would become $$Z_1=Z_2.$$ (77) But from (7.6) then it follows that the ghost self-energy at zero point in the Landau gauge should become zero, i.e. $`\stackrel{~}{b}(0;\xi =0)=0`$. This would be possible only within the nonperturbative treatment of the corresponding ghost self-energy SD equation. The problem is that the perturbative calculation of this renormalization constant within the LA to the above mentioned SD equation suffers from the IR singularities due to exchange of massless gluons. In this case this would mean that the ghost propagator in the Landau gauge does not renormalize at zero point as well as the ghost-gluon vertex and ghost-quark scattering kernel . This would be an example of the nonperturbative finite calculation of the corresponding renormalization constant and deserves furture investigation elsewhere. A few concluding remarks are in order. The SD ghost self-energy equation is complicated nonlinear integral equation relating three various Green’s functions to each other. There are no doubts that this equation may have regular as well as singular solutions even in the LA. However, it is necessary to clearly understand that the disappearance of the ghost-self energy at zero point in the Landau gauge is completely nonperturbative effect. One can not obtain this result calculating only finite number of perturbative corrections. Only summing up infinite number of perturbative diagrams within the corresponding SD equation (i.e. going beyond the perturbation theory), one may hope to achieve this goal. Moreover, it depends heavily on the nature of the approximation scheme. i.e. not any approximation scheme may satisfy this exact result. In other words, the above mentioned possible zero ghost self-energy in the Landau gauge is an example of the nonperturbative finite calculation of the corresponding renormalization constants at zero point since the perturbative renormalization at this point suffers from the IR singularities due to exhange of massless gluons. Precisely for this reason, the perturbative renormalization is always performed at nonzero space-time momentum $`k^2=\mu ^2`$ in order to avoid (not to solve, of course) these difficulties. At the same time, as it seems to us, the nonperturbative renormalization will make it possible to perform finite calculations of the corresponding renormalization constants at zero point without any problems, i. e. only beyond the perturbation theory one is able to renormalize theory at zero point at least in the Landau gauge. This emphasizes once more the special role of the Landau gauge in QCD. ## VIII Conclusions In summary, we have formulated the method how to prove or disprove the LA in gauge theories such as QED and QCD. In QED the summation of the ladder diagrams within the BS integral equation for the quark-photon vertex at zero momentum transfer on account of the corresponding WT identity does not provide an addition constraint on the solution to the quark SD equation itself. In other words, there is no criterion to prove or disprove the use of the LA in QED. In contrast to this, in QCD because of color degrees of freedom the summation of the ladder diagrams within the BS integral equation for the quark-gluon vertex at zero momentum transfer on account of the corresponding ST identity $`does`$ provide an addition constraint on the solution to the quark SD equation itself. Moreover, the solution of the constraint equation (3.14) requires that the full quark propagator should be almost trivial (free-type) one (4.3) or (5.5), i. e. there is $`nonontrivial`$ quark propagator in QCD in the LA. In other words, there is no running quark mass in the LA to QCD as well. This is our main result. It does not depend on how one treats ghost degrees of freedom in the LA to covariant gauge QCD, omitting them by ”hand” in the general constraint equation (3.14) or in more sophisticated fashion by using auxiliarly BS-type integral equations (5.1) and (5.2) there. Let us underline once more, that the existence of the constraint equation (3.14) in the LA to QCD in any gauge is only due to color degrees of freedom and not ghosts. Let us make one thing perfectly clear. The quark SD equation, remainning a nonlinear integral equation even in the LA, may have formally a number of non-trivial, approximate solutions. However, non of them (analytical or numerical) will satisfy the constraint which comes from the exact ST identity, so all results based on these solutions should be reconsidered. In noncovariant gauges ghosts do not contribute from the very beginning, i. e. the ST identity (3.4) becomes of the QED-type apart from the color degrees of freedom. This means that the general constraint equation (3.14) automatically becomes (4.1). Thus from our consideration, one easily can conclude that the quark propagator remains trivial one (4.3) in the LA to noncovariant gauge QCD as well (let us remind that we neither use the explicit expression for the full gluon propagator (2.2) nor specify the full gluon form factor there). The physical reason of the discovered triviality of the quark propagator in the LA is, of course, that the LA in the quark sector ignores the self-interactions of gluons caused by color charges (non-abelian character of QCD). On the other hand, there is no doubt that the LA (especially in the ILA form) is legitimated to use for the ultraviolet (UV) region in QCD because of asymptotic freedom. The problem how to correctly go beyond the LA in order to describethe IR region in QCD becomes inevitably important. It is perfectly clear now that we can not use the LA (in QLA or even in ILA forms (since as it was mentioned above we did not specify the full gluon form factor)) in the entire energy-momentum range in QCD. The ILA itself is not sufficient to approximate the IR region in QCD. The nontrivial generalization of the point-like vertices is needed for this purpose. We have investigated the self-consistency of the LA in QCD. It is almost obvious that any truncation scheme, for example such as planar, $`1/N_c`$ limit, etc maight and should be investigated in the same way. As it is emphasized in our paper, this is important in gauge theories. We didnot discuss the flavor-nonsinglet (but color-singlet) vector vertex to which external gauge fields such as $`W`$ and $`Z`$ bosons could couple. Evidently, the situation for this vertex is completely analogous to that of QED. In other words, contrast to the axial-vector vertex considered in Section 6, here the use of the LA is justified as in QED. First it was explicitly shown in Ref. though in the different way. However, in the light of our result a general question is in order. How is it possible to use nontrivial solutions to the quark SD equation in the LA in the QCD boun-state problems when one knows that it becomes trivial (almost free) in the scattering problems? At long last, the quark SD equation is the same for both problems. At least in one case, namely in the constituent quark model (CQM) the use of the LA is justified for both problems. Indeed, from a QCD theoretical field point of view, apart from other simplifications, the CQM mainly is nothing but an approximation of the full quark by the constituent quark propagator (which coincedes with (5.5) due to replacement $`\stackrel{~}{m}m_q`$, where $`m_q`$ is the constituent quark mass) and the full constituent quark-gluon vertex by point-like one. From our results then it follows that in CQM the quark propagator and quark-gluon vertex is in one-to-one correspondence with each other. Precisely, this self-consistent relation between them explains the success of this calculation scheme despite its simplistic structure. In conclusion, let us make a few technical remarks. In QED expansion in powers of the external momentum makes no great sense since there is no stable bound-states (positronium is unstable). In contrast to this, in QCD the same expansion makes precisely great sense because of the existence of the Goldstone sector there. Many important physical quantities such as scattering lenght, pion charge radius, etc are defined as coefficients of the pion form factor expansion in powers of external momentum. The chiral perturbation theory and its counterpart at the fundamental quark level are also based on these expansions. Thus in QCD zero external momentum transfer has a physical meaning as relating directly to various physical observables. The author would like to thank S.Adler for correspondence and A.T.Filippov and A.A. Slavnov for interesting remarks. He is also grateful to F.Schoberl and W.Lucha for critical discussions and useful remarks on first stage of this investigation.
no-problem/9908/cond-mat9908385.html
ar5iv
text
# SPIN-1/2 PERIODIC NONUNIFORM 𝑋⁢𝑋 CHAINS AND THE SPIN-PEIERLS INSTABILITY (April 27, 1999) ## Abstract Using continued fractions we obtained the exact result for the density of magnon states of the regularly alternating spin-$`\frac{1}{2}`$ $`XX`$ chain with Dzyaloshinskii-Moriya interaction. We examined the stability of the magnetic chain with respect to the spin-Peierls dimerization. Since the discovery of the inorganic spin-Peierls compound CuGeO<sub>3</sub> the interest in the properties of spin-Peierls systems considerably increased . The models that can be examined exactly play an important role in clarifying the generic features of such systems. An example of such a model is the spin-$`\frac{1}{2}`$ $`XX`$ chain that was studied in several papers (note, however, that in the non-adiabatic limit such a spin chain does not permit exact analysis ). The aim of the present study is to examine the influence of an additional Dzyaloshinskii-Moriya coupling on the spin-Peierls dimerization. The presence of such a term for CuGeO<sub>3</sub> was proposed in . The multisublattice spin-$`\frac{1}{2}`$ $`XX`$ chain with the Dzyaloshinskii-Moriya interaction was introduced in . In our study we follow the idea of and compare the total ground state energy of the dimerized and uniform chains. However, in contrast to previous works we use the continued-fraction representation for the one-fermion Green functions that allows a natural extension of the calculations for more complicated lattice distortions having finite period. We consider $`N\mathrm{}`$ spins $`\frac{1}{2}`$ on a circle with the Hamiltonian $`H={\displaystyle \underset{n}{}}\mathrm{\Omega }_ns_n^z+2{\displaystyle \underset{n}{}}I_n\left(s_n^xs_{n+1}^x+s_n^ys_{n+1}^y\right)`$ $`+2{\displaystyle \underset{n}{}}D_n\left(s_n^xs_{n+1}^ys_n^ys_{n+1}^x\right).`$ (1) After the Jordan-Wigner transformation one comes to tight-binding spinless fermions on a circle with complex hopping integrals. We introduce the temperature double-time one-fermion Green functions that yield the density of magnon states $`\rho (E)=\frac{1}{\pi N}_n\text{Im}G_{nn}^{}`$, $`G_{nm}^{}G_{nm}^{}(E\pm iϵ)`$. We further make use of the continued-fraction representation for the required diagonal Green functions $`G_{nn}^{}={\displaystyle \frac{1}{E\pm iϵ\mathrm{\Omega }_n\mathrm{\Delta }_n^{}\mathrm{\Delta }_n^+}},`$ $`\mathrm{\Delta }_n^{}={\displaystyle \frac{I_{n1}^2+D_{n1}^2}{E\pm iϵ\mathrm{\Omega }_{n1}\frac{I_{n2}^2+D_{n2}^2}{E\pm iϵ\mathrm{\Omega }_{n2}_{\mathrm{}}}}},`$ $`\mathrm{\Delta }_n^+={\displaystyle \frac{I_n^2+D_n^2}{E\pm iϵ\mathrm{\Omega }_{n+1}\frac{I_{n+1}^2+D_{n+1}^2}{E\pm iϵ\mathrm{\Omega }_{n+2}_{\mathrm{}}}}}.`$ (2) One immediately notes that for any finite period of varying $`\mathrm{\Omega }_n`$, $`I_n`$, $`D_n`$ the continued fractions $`\mathrm{\Delta }_n^{}`$, $`\mathrm{\Delta }_n^+`$ involved into $`G_{nn}^{}`$ (2) become periodic and thus can be calculated exactly yielding the exact result for the density of states and hence for the thermodynamic quantities of spin model (1). For example, for the periodic chain having period 2 $`\mathrm{\Omega }_1I_1D_1\mathrm{\Omega }_2I_2D_2\mathrm{\Omega }_1I_1D_1\mathrm{\Omega }_2I_2D_2\mathrm{}`$ the described scheme gives $`\rho (E)=\{\begin{array}{cc}0,\hfill & \text{if}Eb_4,b_3Eb_2,b_1E,\hfill \\ \frac{1}{2\pi }\frac{|2E\mathrm{\Omega }_1\mathrm{\Omega }_2|}{\sqrt{(E)}},\hfill & \text{if}b_4<E<b_3,b_2<E<b_1,\hfill \end{array}`$ (5) $`(E)=4_1^2_2^2\left[(E\mathrm{\Omega }_1)(E\mathrm{\Omega }_2)_1^2_2^2\right]^2`$ $`=(Eb_4)(Eb_3)(Eb_2)(Eb_1),`$ $`\left\{b_4b_3b_2b_1\right\}=\{{\displaystyle \frac{1}{2}}(\mathrm{\Omega }_1+\mathrm{\Omega }_2)\pm 𝖻_1,{\displaystyle \frac{1}{2}}(\mathrm{\Omega }_1+\mathrm{\Omega }_2)\pm 𝖻_2\},`$ $`𝖻_1={\displaystyle \frac{1}{2}}\sqrt{\left(\mathrm{\Omega }_1\mathrm{\Omega }_2\right)^2+4\left(|_1|+|_2|\right)^2},`$ $`𝖻_2={\displaystyle \frac{1}{2}}\sqrt{\left(\mathrm{\Omega }_1\mathrm{\Omega }_2\right)^2+4\left(|_1||_2|\right)^2},`$ (6) where $`_n^2=I_n^2+D_n^2`$. To examine the instability of the considered spin chain with respect to dimerization we may assume $`|I_1|=|I|(1+\delta )`$, $`|D_1|=|D|(1+\delta )`$, $`|I_2|=|I|(1\delta )`$, $`|D_2|=|D|(1\delta )`$, $`0\delta 1`$ restricting ourselves to a case of the uniform transverse field $`\mathrm{\Omega }_1=\mathrm{\Omega }_2=\mathrm{\Omega }_0`$. The total energy per site $`(\delta )`$ consists of the magnetic part $`e_0(\delta )`$ that follows from (3) $`e_0(\delta )={\displaystyle \frac{1}{2}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑E\rho (E)|E|={\displaystyle \frac{2||}{\pi }}\text{E}(\psi ,1\delta ^2)|\mathrm{\Omega }_0|\left({\displaystyle \frac{1}{2}}{\displaystyle \frac{\psi }{\pi }}\right),`$ (7) where the effective interspin coupling $`||=\sqrt{I^2+D^2}`$ has been introduced, $`\text{E}(\psi ,a^2)_0^\psi 𝑑\varphi \sqrt{1a^2\mathrm{sin}^2\varphi }`$ is the elliptic integral of the second kind, $`\psi =0`$ if $`2|||\mathrm{\Omega }_0|`$, $`\psi =\text{arcsin}\sqrt{\frac{4^2\mathrm{\Omega }_0^2}{4^2(1\delta ^2)}}`$ if $`2\delta |||\mathrm{\Omega }_0|<2||`$, $`\psi =\frac{\pi }{2}`$ if $`|\mathrm{\Omega }_0|<2\delta ||`$, and the elastic part $`\alpha \delta ^2`$. Besides the trivial solution $`\delta ^{}=0`$ the equation $`\frac{(\delta )}{\delta }=0`$ may have a nonzero one $`\delta ^{}0`$ at moderate and weak fields (i.e. $`|\mathrm{\Omega }_0|<2||`$). This nontrivial $`\delta ^{}`$ comes from the equation that follows from (4) $`{\displaystyle \frac{\pi \alpha }{||}}={\displaystyle \frac{1}{1\delta ^2}}\left(\text{F}(\psi ,1\delta ^2)\text{E}(\psi ,1\delta ^2)\right)`$ (8) where $`\text{F}(\psi ,a^2)_0^\psi 𝑑\varphi /\sqrt{1a^2\mathrm{sin}^2\varphi }`$ is the elliptic integral of the first kind. Consider the case $`\mathrm{\Omega }_0=0`$. Looking for a solution of Eq. (5) that satisfies the inequality $`\delta 1`$ (that is the limit interesting for applications) one observes that the r.h.s. of Eq. (5) can be rewritten approximately as $`_0^1𝑑x/\sqrt{\delta ^2+x^2}`$ ($`x=\mathrm{cos}\varphi `$) that yields $`\delta ^{}\mathrm{exp}\left(\frac{\pi \alpha }{||}\right)`$. post priori we conclude that the small dimerization parameters $`\delta ^{}`$ occur for hard lattices having large values of $`\frac{\alpha }{||}`$. One also notes that the obtained result coincides with the one reported in up to a renormalization of the effective interspin coupling $`|I|||=\sqrt{I^2+D^2}`$. Thus the Dzyaloshinskii-Moriya interaction leads to increasing of the dimerization parameter $`\delta ^{}`$ characterizing the dimerized phase. Consider further the case $`0<|\mathrm{\Omega }_0|<2||`$. Varying $`\delta `$ in the r.h.s. of Eq. (5) from 0 to 1 one calculates a lattice parameter $`\frac{\alpha }{||}`$ for which the taken value of $`\delta `$ realizes an extremum of $`(\delta )`$. One immediately observes that for $`\frac{|\mathrm{\Omega }_0|}{2||}\delta 1`$ the dependence $`\frac{\alpha }{||}`$ versus $`\delta `$ remains as that in the absence of the field, whereas for $`0\delta <\frac{|\mathrm{\Omega }_0|}{2||}`$ the quantity $`\frac{\alpha }{||}`$ starts to decrease. From this one concludes that for hard lattices the field $`\frac{|\mathrm{\Omega }_0|}{2||}=\mathrm{exp}\left(\frac{\pi \alpha }{||}\right)`$ makes the dimerization unstable against the uniform phase. The latter relation tells us that the Dzyaloshinskii-Moriya interaction increases the value of that field. It is generally known that the increasing of the external field leads to a transition from the dimerized phase to the incommensurate phase rather than to the uniform phase. Evidently, the incommensurate phase cannot appear in the presented treatment within the frames of the adopted ansatz for the lattice distortions $`\delta _1\delta _2\delta _1\delta _2\mathrm{}`$, $`\delta _1+\delta _2=0`$. To clarify a possibility of more complicated distortions the chains with longer periods should be examined. Alternatively, we may also assume different dependences on $`\delta `$ for the isotropic coupling and the Dzyaloshinskii-Moriya coupling, for example, $`|I_1|=|I|(1+\delta )`$, $`|D_1|=|D|`$, $`|I_2|=|I|(1\delta )`$, $`|D_2|=|D|`$. Supposing that $`\delta 1`$ after simple rescaling arguments one finds that the dimerization parameter $`\delta ^{}\frac{^2}{I^2}\mathrm{exp}\left(\frac{\pi }{||}\frac{^4}{I^4}\alpha \right)`$. Thus, in such a case the Dzyaloshinskii-Moriya interaction leads to a decreasing of the dimerization parameter characterizing the dimerized phase. The value of the field which destroys dimerization $`\frac{|\mathrm{\Omega }_0|}{2||}=\mathrm{exp}\left(\frac{\pi }{||}\frac{^4}{I^4}\alpha \right)`$ decreases as well. To conclude, we have analysed a stability of the spin-$`\frac{1}{2}`$ transverse $`XX`$ chain with respect to dimerization in the presence of the Dzyaloshinskii-Moriya interaction calculating for this purpose with the help of continued fractions the ground state energy for an arbitrary value of the dimerization parameter. Assuming that the ratio of the Dzyaloshinskii-Moriya coupling to the isotropic coupling does not depend on the dimerization parameter we have found that the Dzyaloshinskii-Moriya interaction leads to an increasing of the effective interspin coupling and thus to some quantitative changes, i.e. to an increasing of the value of the dimerization parameter which characterizes the dimerized phase and the value of the field which destroys the dimerized phase. In the other limiting case when the Dzyaloshinskii-Moriya coupling does not depend on the dimerization parameter it has an opposite effect leading to a decreasing of the value of the dimerization parameter and the value of the field which destroys dimerization. The obtained results are in agreement with some earlier studies of the thermodynamic properties of spin-$`\frac{1}{2}`$ transverse $`XX`$ chains with Dzyaloshinskii-Moriya interaction . Finally, it is known that the Dzyaloshinskii-Moriya interaction may lead to drastical changes in spin correlations and a study of a relation of these changes to the spin-Peierls instability seems to be an interesting issue. The present study was partly supported by the DFG (projects 436 UKR 17/20/98 and Ri 615/6-1). O. D. acknowledges the kind hospitality of the Magdeburg University in the spring of 1999 when the main part of the paper was done. He is also indebted to Mrs. Olga Syska for continuous financial support.
no-problem/9908/cond-mat9908288.html
ar5iv
text
# Phase diagram for morphological transitions of wetting films on chemically structured substrates ## I Introduction At present various experimental techniques allow one to provide substrates with well-defined, microscopically small geometrical or chemical structures or combinations thereof (see, e.g., Ref. ). These structured substrates can serve as devices to guide and process tiny amounts of liquids. Such “microfluidics” systems are promising for future applications, e.g., in chemistry and biology . Although the performance of these systems ultimately depends on their dynamic properties, a thorough understanding of the corresponding structural properties in thermal equilibrium is an important prerequisite . In this spirit as a paradigmatic case we determine the equilibrium structures of liquidlike wetting films forming at the interface between a bulk vapor phase and a flat solid substrate containing a single chemical stripe (Fig. 1). Our aim is to explore the entire phase diagram for the emerging lateral fluid structures as function of temperature, undersaturation or pressure, and stripe width. In Ref. it has been shown that a thin liquidlike film adsorbed on such a substrate (Fig. 1) can undergo an interesting morphological phase transition (Figs. 2 and 3) of the shape $`l(x)`$ of the liquid-vapor interface, depending on the structures of the effective interface potentials $`\mathrm{\Lambda }_\pm (l)`$ characterizing the corresponding *homogeneous* substrates “$`+`$” and “$``$” (see, e.g., Fig. 4). In accordance with Ref. we study a substrate designed such that the effective interface potential $`\mathrm{\Lambda }_+(l)`$ of the stripe part exhibits two competing local minima (“I” near and “III” further away from the substrate surface) whereas the asymptotic film thickness $`l_{}`$ far from the stripe is given by the single global minimum (“II”) of the interface potential $`\mathrm{\Lambda }_{}(l)`$ of the embedding substrate (see Fig. 4). In Ref. the phase diagram for this morphological phase transition has been determined, at constant temperature $`T`$, as function of the stripe width $`a`$ and of the undersaturation $`\mathrm{\Delta }\mu =\mu _0\mu 0`$ where $`\mu _0(T)`$ is the chemical potential at liquid-vapor coexistence. The purpose of the present study is to determine how this phase diagram for the morphological phase transition within the $`a`$-$`\mathrm{\Delta }\mu `$ plane evolves as function of temperature. This additional information is important in order to be able to put the morphological phase transitions into the context of $`T`$-$`\mathrm{\Delta }\mu `$ wetting phase diagrams, including the prewetting line, on homogeneous substrates . We emphasize that experimental techniques such as reflection interference contrast microscopy (RICM) and atomic force microscopy (AFM) in tapping mode are capable of imaging the liquidlike structures under consideration down to the nm scale. Moreover, due to their relatively small spatial extensions the structures discussed here are well accessible by Monte Carlo simulations as another means for testing the theoretical predictions. For these envisaged studies the knowledge of the full topology of the phase diagram as function of $`T`$, $`\mathrm{\Delta }\mu `$, and $`a`$ is very important. ## II Theoretical model We use the following simple interface displacement model within which the equilibrium interface profile $`\overline{l}(x)`$ minimizes the functional $$\mathrm{\Omega }_𝒮[l(x)]=_A𝑑x𝑑y\mathrm{\Lambda }(x,l(x))+\sigma _{lg}_A𝑑x𝑑y\left(\frac{dl(x)}{dx}\right)^2$$ (1) where $`A=L_xL_y`$ is the area of the flat substrate surface located at $`z=0`$. The first part takes into account the effective interaction between the liquid-vapor interface and the substrate via the laterally varying effective interface potential $`\mathrm{\Lambda }(x,l)=\mathrm{\Delta }\mathrm{\Omega }_bl+\omega (x,l)`$. The second part in Eq. (1) is the leading-order term of a gradient expansion of the surface free energy associated with the deviation of the liquid-vapor interface profile from its flat configuration; $`\sigma _{lg}`$ is the surface tension. $`\mathrm{\Delta }\mathrm{\Omega }_b=\mathrm{\Delta }\mu (\rho _l\rho _g)+𝒪((\mathrm{\Delta }\mu )^2)`$ is the difference of the bulk free energy densities of the liquid and the vapor phase with number densities $`\rho _l`$ and $`\rho _g`$, respectively. Alternatively, in terms of the bulk pressure $`p`$, $`\mathrm{\Delta }\mathrm{\Omega }_b`$ is given approximately by the Gibbs-Thomson expression $`k_BT\rho _l\mathrm{ln}(p_{sat}/p)`$ where $`p_{sat}`$ is the saturated vapor pressure. At liquid-vapor coexistence $`\mathrm{\Lambda }(x,l)`$ is given by $`\omega (x,l)`$ which is of the form $`_{i2}a_i(x)/l^i`$. For reasons of simplicity we assume that the lateral variation of $`\omega (x,l)`$ is steplike: $`\omega (x,l)=\mathrm{\Theta }(|x|a/2)\omega _{}(l)+\mathrm{\Theta }(a/2|x|)\omega _+(l)`$ with the effective interface potential $`\omega _+`$ ($`\omega _{}`$) of a *homogeneous* substrate composed of “$`+`$” (“$``$”) particles. In Refs. and we have shown that the functional in Eq. (1) – despite its simplicity – allows one to determine reliably the morphology of liquidlike films on structured substrates, and we have presented a derivation of this square-gradient functional from a microscopic density functional theory. We also found (see Fig. 11 in Ref. ) that the aforementioned steplike variation of $`\omega (x,l)`$ as opposed to its actual smooth lateral variation represents a rather reliable approximation. The form of $`\omega (x,l)`$ follows from considering as the interaction potential between the fluid particles a Lennard-Jones potential with a depth $`ϵ`$ and an effective particle diameter $`\sigma `$. Moreover, also the interaction between the fluid and the substrate particles is modeled by a Lennard-Jones potential (compare Ref. ). Since the substrate is translationally invariant in the $`y`$ direction the effective interface potential depends only on $`x`$, so that $`\overline{l}(x)`$ is also translationally invariant in the $`y`$ direction. The interaction potential parameters of the stripe part (“$`+`$”) are the same as for the “$`+`$” substrate described in Fig. 11 in Ref. , and for the embedding “$``$” substrate part we use the parameters for the “$`+`$” substrate part in Fig. 7 in Ref. multiplied by $`0.9102`$. The temperature dependence of the effective interface potentials $`\mathrm{\Lambda }_\pm (l)`$ is determined by systematically keeping track of the temperature dependence of the bulk liquid and vapor densities $`\rho _l`$ and $`\rho _g`$ which enter into $`\mathrm{\Lambda }_\pm (l)`$ as discussed in Sec. II in Ref. . For the case studied here the temperature dependence of $`\omega _+(l)`$ for the stripe part is such that a homogeneous “$`+`$” substrate exhibits a first-order wetting transition at $`T_w^{}=k_BT_w/ϵ1.102`$ and prewetting transitions along the prewetting line $`\mathrm{\Delta }\mu _{pre}(T)`$, $`T>T_w`$ and $`\mathrm{\Delta }\mu >0`$, that extends into the vapor phase region of the phase diagram (thick dashed line denoted as “P” in Fig. 2). $`\omega _{}(l)`$ for the embedding substrate is also temperature dependent such that the asymptotic film thickness $`l_{}`$ grows with increasing $`T`$ and shrinks with increasing $`\mathrm{\Delta }\mu `$. As it turns out, however, the exact temperature dependence and the magnitude of $`l_{}`$ are not important for the qualitative features of the phase diagram for the morphological transitions. The systems under consideration here differ from those studied in Ref. in two important aspects. First, here the liquidlike films are so thin that they are completely under the influence of the substrate potential which determines, inter alia, the effective interface potential. Secondly, we consider a grand canonical ensemble without a volume constraint for the liquid phase so that the liquidlike films are not subject to instabilities along the $`y`$ direction as observed in, e.g., Ref. . Using a numerical relaxation technique we solve the two-point boundary value problem for the Euler-Lagrange equation $$\sigma _{lg}\frac{d^2l(x)}{dx^2}=\mathrm{\Delta }\mathrm{\Omega }_b+\frac{\omega (x,l)}{l}|_{l=l(x)}\text{with}l(x\pm \mathrm{})=l_{}$$ (2) which follows from Eq. (1) by functional differentiation with respect to $`l(x)`$. Equation (2) exhibits the structure of a one-dimensional classical mechanical equation of motion in a time-dependent external potential. With the stripe width $`a`$ and the temperature $`T`$ fixed, for a wide range of undersaturations there are two solutions of Eq. (2) which correspond to local minima of the functional $`\mathrm{\Omega }_𝒮`$. One of the solutions is closely bound to the stripe whereas the other solution is further away or even repelled from the stripe as described in Ref. . Figures 47 show pertinent examples for the effective potentials $`\mathrm{\Lambda }_\pm `$ of the two substrate materials and the two appertaining solutions of Eq. (2) at three different points of the phase diagram: Figs. 4, 5, 6, and 7 correspond to the points $`(T,\mathrm{\Delta }\mu )`$ indicated by “a”, “b”, “c”, and “d”, respectively, in the phase diagram shown in Figs. 2 and 3. The equilibrium solution $`\overline{l}(x)`$ for the profile corresponds to the global minimum of $`\mathrm{\Omega }_𝒮`$, and the other solution is metastable. At a certain value $`\mathrm{\Delta }\mu _t`$, for which $`\mathrm{\Omega }_𝒮`$ has the same value for both solutions, a phase transition from one interfacial configuration to the other takes place. This transition is first order because the derivatives of $`\mathrm{\Omega }_𝒮(\overline{l}(x))`$ with respect to $`\mathrm{\Delta }\mu `$ are discontinuous. As already shown in Ref. , in Eq. (1) the two contributions to $`\mathrm{\Omega }_𝒮`$ compete with each other in minimizing the whole functional. Depending on the special choices for $`a`$, $`T`$, and $`\mathrm{\Delta }\mu `$ the cost in free energy for increasing the liquid-vapor interface area is overcompensated by the gain in free energy which follows from occupying the deeper minimum (I) of $`\mathrm{\Lambda }_+`$ near the wall, so that the bound configuration has the lower free energy. In the opposite case this gain in free energy is too small to compensate the cost in free energy from the increased area of the liquid-vapor interface, leading to a less bound or even repelled interface profile. ## III Discussion of the phase diagram In Fig. 2 the lines $`\mathrm{\Delta }\mu _t(T;a)=\mu _0(T)\mu _t(T;a)`$ of phase coexistence between the bound and the repelled configuration for different stripe widths $`a`$ are presented. These lines are vertical cross-sections of the full phase diagram shown in Fig. 3. For a given stripe width $`a`$ the bound (repelled) solution is stable for $`\mathrm{\Delta }\mu <\mathrm{\Delta }\mu _t(T;a)`$ ($`\mathrm{\Delta }\mu >\mathrm{\Delta }\mu _t(T;a)`$), i.e., below (above) the corresponding line of phase coexistence. The triangles indicate the intersections $`T_0(a)`$ between the coexistence lines $`\mathrm{\Delta }\mu _t(T;a)`$ and the bulk liquid-vapor coexistence line $`\mathrm{\Delta }\mu =0`$. At $`T_0(a)`$ the morphological phase transition occurs at liquid-vapor coexistence, forming the line denoted as $`\alpha `$ in Fig. 3. Upon increasing the stripe width $`a`$ the lines of coexistence are shifted towards the prewetting line “P” without touching or crossing it for any finite value of $`a`$. (This means that for large $`a`$ the lines of coexistence do not end and reappear at the prewetting line such as, e.g., the prefilling transition lines in a wedge-shaped groove for large opening angles of the wedge .) For constant undersaturation $`\mathrm{\Delta }\mu `$ we find $`T_{pre}(\mathrm{\Delta }\mu )T_t(\mathrm{\Delta }\mu ;a\mathrm{})a^1`$ for the difference between the transition temperatures $`T_t(\mathrm{\Delta }\mu ;a)`$ for the morphological and the prewetting transition. In the limit $`a\mathrm{}`$ the morphological phase transition tallies with the prewetting transition on the homogeneous “$`+`$” substrate. On the mean-field level considered hitherto each coexistence curve $`\mathrm{\Delta }\mu _t(T;a)`$ is a line of first-order transitions ending in a critical point $`(T_c(a),\mathrm{\Delta }\mu _c(a))`$ (denoted as full diamonds in Fig. 2 and forming the line denoted as $`\gamma `$ in Fig. 3) such that at $`T>T_c(a)`$ or $`\mathrm{\Delta }\mu <\mathrm{\Delta }\mu _c(a)`$ for the given stripe width $`a`$ there is no morphological transition but a smooth variation from the bound to the repelled solution. As demonstrated by Fig. 2 and its inset the positions of the critical points exhibit a nontrivial and nonmonotonous dependence on the stripe width $`a`$; there is no simple criterion for the effective interface potential $`\mathrm{\Lambda }_\pm `$ which allows one to predict the corresponding line of critical points (see the line $`\gamma `$ in Fig. 3) as function of $`a`$. As already pointed out in detail in Ref. interface fluctuations along the $`y`$ direction in this effectively one-dimensional stripe configuration actually smear out the sharp first-order morphological phase transitions and thus eliminate the critical points . In Fig. 2 for $`a=5\sigma `$ the width of this fluctuation-induced smooth transition region is indicated by the thin dashed-dotted lines. For $`a10\sigma `$ these fluctuation effects are already negligibly small, apart from the close vicinity of the critical points which are still eliminated for any finite $`a`$. However, in the limit $`a\mathrm{}`$ the coexistence lines $`\mathrm{\Delta }\mu _t(T;a)`$ merge with the prewetting line “P” (see Fig. 3) associated with the homogeneous “$`+`$” substrate, which does have a genuine critical point “C” beyond mean field theory. The three thermodynamic states denoted as “a”, “b”, and “c” in Figs. 2 and 3 lead to the interface configurations shown in Figs. 4, 5, and 6, respectively. The corresponding effective interface potentials $`\mathrm{\Lambda }_\pm (l)`$ differ with respect to the number and the relative positions of their local minima. For temperatures $`T<T_{pre}`$ and small undersaturations $`\mathrm{\Delta }\mu `$, $`\mathrm{\Lambda }_+(l)`$ exhibits two local minima (I and III); an example for this case is shown in Fig. 4 appertaining to the thermodynamic state denoted as “a” in Fig. 2. This case is analogous to that described in Ref. . At the thin dotted line $`\mathrm{\Delta }\mu _i(T)`$ denoted as “i” in Fig. 2 the minimum III of $`\mathrm{\Lambda }_+`$ far from the wall and the maximum of $`\mathrm{\Lambda }_+`$ merge, forming a saddle point; $`\mathrm{\Delta }\mu _i(T)`$ ends at the prewetting critical point “C”. For $`\mathrm{\Delta }\mu \mathrm{\Delta }\mu _i(T)`$ there is only one local minimum (I) of $`\mathrm{\Lambda }_+`$. An example for this latter situation is shown in Fig. 5 corresponding to the thermodynamic state denoted as “b” in Fig. 2. Surprisingly the coexistence lines $`\mathrm{\Delta }\mu _t(T;a)`$ for the morphological phase transition extend below this line “i”. This shows that the presence of an energy barrier and of a second local minimum of $`\mathrm{\Lambda }_+`$ is not a necessary condition for the occurrence of the morphological phase transition described here and in Ref. . The thin dotted line $`\mathrm{\Delta }\mu _{iii}(T)`$ denoted as “iii” is the line at which, analogously, the minimum I near the wall and the maximum of $`\mathrm{\Lambda }_+`$ merge, forming a saddle point. Both lines “i” and “iii” meet and end at the critical point “C” of the prewetting transition of a homogeneous “$`+`$” substrate where both minima I and III and the maximum of $`\mathrm{\Lambda }_+`$ merge, leaving a single minimum. At the thin short-dashed line $`\mathrm{\Delta }\mu _{ii}(T)`$ denoted as “ii” in Fig. 2 the positions of the second minimum III of $`\mathrm{\Lambda }_+`$ and of the minimum II of $`\mathrm{\Lambda }_{}`$ coincide. Along this line one of the solutions of Eq. (2) is completely flat: due to $`d\mathrm{\Lambda }_\pm /dl(l=l_{})=0`$ Eq. (2) leads to the trivial solution $`l(x)l_{}`$ with $`d^2l(x)/dx^20`$. An example for this case is shown in Fig. 6 which corresponds to the thermodynamic state denoted as “c” in Fig. 2. The line “ii” separates the region where the repelled solution is bent towards the stripe ($`\mathrm{\Delta }\mu <\mathrm{\Delta }\mu _{ii}(T)`$) from that where it is bent away from the stripe ($`\mathrm{\Delta }\mu >\mathrm{\Delta }\mu _{ii}(T)`$). We note that the fact that one of the solutions is completely flat is an artifact of our simplifying model assumption that $`\omega (x,l)`$ varies steplike as function of $`x`$. In a more realistic model with a smooth lateral variation of the effective interface potential (as used in Ref. ) the corresponding solution would be almost flat with a small curvature. In the system considered here the asymptotic film thickness $`l_{}`$, i.e., the position of minimum II, is always larger than the position of minimum I. However, for low temperatures there is a line at which the minimum II of $`\mathrm{\Lambda }_{}`$ and the *maximum* of $`\mathrm{\Lambda }_+`$ coincide. This occurs along the thin long-dashed line $`\mathrm{\Delta }\mu _{iv}(T)`$ denoted as “iv” in Fig. 2. This line is the continuation of line “ii” for lower temperatures $`T`$ starting at the point where the line “ii” meets the line “i” tangentially. Along this line, too, the repelled solution of Eq. (2) is constant, i.e., $`l(x)l_{}`$. As the line “ii”, also line “iv” separates two regions with different curvature behavior of the repelled solution. For temperatures $`T>T_{pre}`$, i.e., on the right side of the prewetting line P, the minimum III of $`\mathrm{\Lambda }_+`$ is deeper than the minimum I. The difference between the asymptotic film thickness $`l_{}`$ corresponding to the minimum II and the deeper minimum III is also smaller than that between $`l_{}`$ and the local minimum I near the wall. Therefore both contributions to $`\mathrm{\Omega }_𝒮`$ are larger in the case that the interface follows minimum I as compared to the case that it follows minimum III so that for $`T>T_{pre}`$ the bound solution is always metastable. This situation is shown in Fig. 7 which corresponds to the thermodynamic state denoted as “d” in Fig. 2. If $`TT_{iii}(\mathrm{\Delta }\mu )`$ $`\mathrm{\Lambda }_+`$ exhibits only one local minimum (III) and therefore there is only one solution of Eq. (2), namely the repelled one. ## IV Conclusion Within an interface displacement model based on a microscopic density functional theory we have calculated the entire phase diagram (Figs. 2 and 3) of morphological phase transitions of wetting films on a substrate with a chemical stripe (Fig. 1). This phase diagram and the corresponding equilibrium interface profiles shown in Figs. 47 elucidate the dependence of the morphological phase transitions on the thermodynamic variables temperature and undersaturation (or, equivalently, pressure) as well as on the stripe width. The morphological phase transitions have been put into the context of the prewetting transitions occuring on a homogeneous substrate formed by particles of the stripe material. ###### Acknowledgements. We gratefully acknowledge financial support by the German Science Foundation within the Special Research Initiative *Wetting and Structure Formation at Interfaces*.
no-problem/9908/cond-mat9908291.html
ar5iv
text
# Ising cubes with enhanced surface couplings ## I Introduction In the thermodynamic limit, critical phenomena may occur not only in the bulk of a system, but also at its surfaces, edges and corners. To be specific, let us consider Ising magnets with short-range interactions. Then, there are two typical scenarios: (a) bulk, $`m_b`$, surface, $`m_1`$, edge, $`m_2`$, and corner, $`m_3`$, magnetizations may order at the same temperature (”ordinary transition”), but with different power–laws, and (b) surface, edge and corner magnetizations may order simultaneously first (”surface transition”) due to enhanced, strong surface couplings, followed by ordering of the bulk magnetization at the lower bulk transition temperature (”extraordinary transition”). Surface singularities at ordinary and surface transitions have been studied extensively, theoretically as well as experimentally . Most of the rather few studies on edge critical behavior dealt with the ordinary transition . Only very recently, edge criticality both at the surface transition and at the normal transition has been investigated. Similarly, corner criticality in three-dimensional Ising systems has been analysed, to our knowledge, only at the ordinary transition, applying mean-field theory and Monte Carlo simulations . However, that case deserves to be studied at the surface transition as well for various reasons. For magnetic properties of nanostructured materials, corners are expected to play an important role . In addition, magnetic couplings may be enhanced at surfaces, especially at step-edges and corners . Last, but not least, the problem is of genuine theoretical interest. At the surface transition, the critical fluctuations are essentially two–dimensional. Edges are local perturbations acting then like defect lines in two-dimensional Ising models leading to interesting non–universal critical phenomena . Accordingly, corners, say, of an Ising cube may be interpreted as intersection points of three defect lines. At such points, one also expects intriguing non-universal behavior of local quantities, such as the corner magnetization, following exact analytical work on two–dimensional Ising systems with intersecting defect lines . ## II Model, method and results We study nearest–neighbor Ising models on simple cubic lattices with $`L\times M\times N`$ spins (usually we shall consider Ising cubes, i.e. $`L=M=N`$) and ferromagnetic interactions. The Hamiltonian may be written in the form $``$ $`=`$ $`{\displaystyle \underset{bulk}{}}J_bS_{xyz}S_{x^{}y^{}z^{}}{\displaystyle \underset{surface}{}}J_sS_{xyz}S_{x^{}y^{}z^{}}`$ (2) $`{\displaystyle \underset{edgesurface}{}}J_{es}S_{xyz}S_{x^{}y^{}z^{}}{\displaystyle \underset{edge}{}}J_eS_{xyz}S_{x^{}y^{}z^{}}`$ where the sums run over bonds between neighboring spins, $`S_{xyz}=\pm 1`$, with coupling constants to be specified below; $`x(y,z)`$ going from 1 to $`L(M,N)`$. Free boundary conditions hold for the spins in the surface planes. The pairs of neighboring spins in the Hamiltonian (1) are located either on edge sites with the edge coupling $`J_e`$, on edge and surface sites coupled by $`J_{es}`$, on surface sites with the interaction $`J_s`$, or on sites with at least one of the spins in the interior of the system interacting with the bulk coupling $`J_b`$. We refrained from assigning another coupling strength to pairs of spins on corner and edge sites, which is taken to be equal to $`J_e`$. To study the behavior at the surface transition, $`T_s`$, we chose $`J_s=2J_b`$, where $`k_BT_s/J_b4.975`$ , while the bulk transition, $`T_c`$, occurs at $`k_BT_c/J_b4.5115`$ . The effect of the edge couplings was studied by considering the three cases (i) of equal surface couplings, i.e. $`J_e=J_{es}=J_s`$, (ii) of reduced edge couplings, especially $`J_{es}=J_s`$, $`J_e=J_b`$, and (iii) of reduced edge–surface couplings, especially $`J_e=J_s`$, $`J_{es}=J_b`$. The size of the Ising cubes wih $`L^3`$ spins ranged from $`L=5`$ to $`L=80`$. In the Monte Carlo simulations, we used the efficient single–cluster–flip algorithm. Thermal averages were obtained from an ensemble of at least $`10^2`$ realizations, using different random numbers. In each realization, several $`10^4`$ clusters were taken into account, after equilibration. The quantity of main interest is the corner magnetization, or more general, the local magnetization $`m_l(x,y,z)`$ at site $`(xyz)`$. $`m_l(x,y,z)`$ may be defined by the correlation function $$m_l(x,y,z)=\sqrt{<S_{xyz}S_{x^{}y^{}z^{}}>}$$ (3) where $`(xyz)`$ and $`(x^{}y^{}z^{})`$ are topologically equivalent sites with maximal separation distance; brackets denote the thermal average. In the thermodynamic limit, $`L\mathrm{}`$, one recovers the standard definition of the local magnetization $`m(x,y,z)=<S_{xyz}>`$. $`m_l(x,y,z)`$ approaches closely $`m(x,y,z)`$ provided the separation distance between the two equivalent spins is large compared to the correlation length. Certainly, finite–size effects are most severe near criticality, as usual. The deviation of $`m_l`$ from $`m`$ may be monitored by varying the size of the cubes, $`L`$, and by considering the correlation function between spins on equivalent sites with different separation distances (for instance, two corner spins may be connected either by an edge, by a surface diagonal, or by the bulk diagonal). In Fig. 1 the intriguing profile of the local magnetization at the surface, $`m_l(x,y,1)`$, of a $`40^3`$ Ising cube with equal surface couplings, case (i), is depicted, at $`k_BT/J_b=4.9`$, i.e. $`T0.985T_s`$. The non–monotonic behavior along pathes from the edges or corners towards the center of the surface reflects the influence of bulk spins, as has been discussed before . Crossover to monotonicity of the profile, with the largest magnetization at the corners, is expected to occur even closer to $`T_s`$. At lower temperatures, roughly $`T<T_c`$, the profile is monotonic as well, with the smallest magnetization at the corners due to the different coordination numbers at corners, edges, and surfaces. Near the surface transition, $`T_s`$, the corner magnetization, say, $`m_3=m_l(1,1,1)`$, is expected to vanish, in the thermodynamic limit, as $`m_3t^{\beta _3}`$, where $`t`$ is the reduced temperature $`t=|TT_s|/T_s`$. To estimate $`\beta _3`$, we consider the effective exponent $`\beta _{eff}(t)=d\mathrm{ln}m_3/d\mathrm{ln}t`$. When analysing the Monte Carlo data, the derivative is replaced by a difference at discrete temperatures. As $`t0`$, $`\beta _{eff}`$ approaches $`\beta _3`$, provided finite–size effects can be neglected. The temperature dependence of the effective exponent $`\beta _{eff}`$ for the three different sets of couplings, (i), (ii), and (iii), is shown in Fig. 2, displaying only data which were checked to be unaffected by finite–size effects. Error bars stem from the ensemble averaging performed to determine $`m_3`$. The resulting estimates for the asymptotic critical exponent $`\beta _3`$ are (i) $`0.06\pm 0.01`$ at $`J_e=J_{es}=J_s=2J_b`$, (ii) $`0.14\pm 0.015`$ at $`J_{es}=J_s=2J_b`$ and $`J_e=J_b`$, and (iii) $`0.26\pm 0.02`$ at $`J_e=J_s=2J_b`$ and $`J_{es}=J_b`$. In addition, we estimated $`\beta _3`$ at the ordinary transition, $`J_e=J_{es}=J_s=J_b`$ to be $`\beta _3=1.77\pm 0.05`$, confirming and refining our previous estimate based on computing the corner magnetization from metastable states .– Error bars are inferred from ”reasonable” extrapolations of the effective exponent, see Fig. 2. To explain the Monte Carlo findings on $`\beta _3`$, note that the critical fluctuations at the surface transition are essentially two–dimensional and that corners are intersection points of the edges. Now, as had been shown before , at the surface transition edges act as ladder– or chain–type defect lines . The critical exponent $`\beta _2`$ of the edge magnetization is non–universal (being non–trivial even in case (i) of equal surface couplings, due to the coupling to bulk spins), varying with the edge, $`J_e`$, and edge–surface, $`J_{es}`$, couplings. To a given set of interactions $`J_e`$ and $`J_{es}`$, one may assign roughly an effective defect coupling of ladder– or chain–type, $`J_d^{eff}`$, yielding the same critical exponent for the defect magnetization in the two–dimensional Ising model, $`\beta _l`$, and for the edge magnetization at the surface transition of the three–dimensional Ising model, $`\beta _2`$. Specificly, for a ladder–type defect, the critical exponent of the magnetization in the ladder rows, $`\beta _l`$, is given by $$\beta _l=2\mathrm{arctan}^2(\kappa _l^1)/\pi ^2$$ (4) with $`\kappa _l=\mathrm{tanh}(J_l/(k_BT_{2d}))/\mathrm{tanh}(J/(k_BT_{2d}))`$, where $`T_{2d}`$ is the transition temperature. Comparing $`\beta _2`$ and $`\beta _l`$, one may interpret the defect coupling $`J_l`$ of the two–dimensional model as the desired effective coupling $`J_d^{eff}`$ ($`J`$ is the coupling constant away from the defect line, corresponding to $`J_s`$ in the three–dimensional systems). Following this analogy, the critical exponent of the corner magnetization, $`\beta _3`$, can be related to that of the magnetization at the intersection of three defect lines in the two–dimensional Ising model, $`\beta _i`$, with effective defect couplings, $`J_d^{eff}`$. Indeed, in the two–dimensional Ising model, the value of $`\beta _i`$ has been calculated exactly for three intersecting ladder defects by Henkel et al. , showing a non–universal behavior, with $`\beta _i`$ depending on the strength of the defect couplings, $`J_l`$. If those couplings are weaker than in the rest of the system, then $`\beta _i`$ will increase with decreasing $`J_l(<J)`$, $`\beta _i>1/8`$, 1/8 being the well–known Onsager value in the isotropic two–dimensional Ising model. In turn, if the defect couplings get stronger, then $`\beta _i`$ will get smaller. The concrete expression for $`\beta _i`$ is quite lengthy and will not be reproduced here, but it can be evaluated in a straightforward way. The effective ladder-type defect couplings $`J_d^{eff}`$ in the three cases we considered are (i) $`J_d^{eff}1.22J_s`$ corresponding to $`\beta _20.095`$ in the case of equal surface couplings, i.e. an effective enhancement of the couplings at the edges due to the influence of bulk spins, (ii) $`J_d^{eff}0.99J_s`$ corresponding to $`\beta _20.127`$ for weakened edge couplings, i.e. the enhancement is now approximately compensated by the weakening of $`J_e`$, and (iii) $`J_d^{eff}0.74J_s`$ corresponding to $`\beta _20.176`$ for weakened edge–surface couplings, overcompensating the enhancement by the reduction in $`J_{es}`$ (note that the values of $`\beta _2`$ differ significantly from those of $`\beta _3`$). Using these estimates of $`J_d^{eff}`$, one obtains from the exact expression for the two–dimensional Ising model with three intersecting ladder defects of those strengths the following values for $`\beta _i`$ (i) 0.082 , (ii) 0.128, and (iii) 0.21, in satisfactory agreement with the Monte Carlo findings on $`\beta _3`$. Of course, a more refined analysis had to take into account, e.g., the rather complicated (see also the non–monotonic profile in Fig. 1) nature of the edge as a simultaneously ladder– and chain–type defect line as well as the effect of the bulk spin next to the corner on the corner magnetization. Indeed, the good agreement between $`\beta _3`$ and $`\beta _i`$ in case (ii) may be related to the fact that the chain–like character is rather weak in that situation. The bulk spin is expected to strengthen the effective coupling at the corner especially in case (i), giving rise to the reduction in $`\beta _3`$ as compared to $`\beta _i`$. Certainly, bulk properties will become critical only at the extraordinary transition, at $`k_BT_c/J_b4.5115`$. For instance, the specific heat $`C`$ is expected to diverge there, in the thermodynamic limit. For finite, $`L^3`$ Ising cubes, one observes that a maximum in $`C`$ near $`T_c`$ shows up only for systems with at least a few thousands spins, getting more pronounced as the system size increases (we studied case (i) with equal surface couplings and $`J_s=2J_b`$). On the other hand, the maximum in $`C`$ near the surface transition, $`T_s`$, dominates for small cubes, becoming more and more suppressed as one increases the size, $`L`$. For cubes of moderate size, say $`15<L<60`$, the temperature dependence of the specific heat is characterised by an easily detectable two–peak structure, with maxima close to $`T_c`$ and $`T_s`$.-The height of the two peaks may be easily varied by replacing the Ising cubes by slabs. In summary, the corner magnetization at the surface transition of Ising cubes has been found to display non-universal critical behavior, with the critical exponent $`\beta _3`$ of the corner magnetization (being distinct from the corresponding edge exponent $`\beta _2`$) depending on the strength of the edge and edge–surface couplings. The concrete value of $`\beta _3`$ may be approximated rather well from the exactly known value of the critical exponent of the magnetization at the intersection point of three defect lines in the two–dimensional Ising model by estimating effective defect couplings from the edge critical behavior. ###### Acknowledgements. It is a pleasure to thank M. Henkel and I. Peschel for very helpful suggestions and discussions.
no-problem/9908/cond-mat9908321.html
ar5iv
text
# Novel surface state in a class of incommensurate systems ## I Introduction It is well known that a surface field can give rise to wetting phenomena and also that enhanced interactions near a surface can give rise to surface order without bulk order . Nakanishi and Fisher have given a unified picture of wetting and surface ordering at the phenomenological (Landau-theory) level; these effects require that surface terms be added to the bulk free energy. In this article, we report an entirely new surface effect which should occur in a particular class of incommensurate systems. We find that surface states exist and are energetically favored by the mere presence of the surface, without surface terms like those considered in . Candidate physical systems for observing these states include highly chiral cholesterics in electric or magnetic field, where a bulk undulating phase was recently predicted to occur . This phase is an undulating structure in which the amount of orientational order varies periodically in conjunction with an oscillation of the direction of the local optic axis. It is expected to occur under appropriate conditions of temperature and a strong aligning electric or magnetic field. As discussed in section III C of reference , the order parameters for the modulated state are the amplitudes of the harmonics. The free energy that results is identical to that of Landau models in which the coefficient of the gradient-squared elastic terms is negative, necessitating the inclusion of terms quadratic in second derivatives. When the coefficient of the gradient-squared term vanishes, a Lifshitz point occurs in the phase diagram . Therefore other candidates include Lifshitz-point systems such as the magnetic material MnP , and Langmuir monolayers and diblock copolymers with modulated phases. Our Landau model gives a bulk temperature-field phase diagram with a closed loop separating the modulated phase (favored inside) from the homogeneous phase. The surprise is that the mere existence of the surface produces a surface state which is energetically favored within a second closed loop well outside the first. Outside the second loop, the surface state exists but it is metastable (the equilibrium solution is simply the homogeneous bulk state). The order parameter in the surface state is not a small perturbation to the bulk order parameter. The width of the state has no pronounced temperature dependence; in particular, it does not diverge. In the presence of a surface field (coupling linearly to the order parameter), the surface phase may still occur but the line of surface transitions no longer forms a closed loop. ¿From the above and other evidence, our surface states are very different from the states considered in . This paper is organized as follows. Section II presents the effective Landau-Ginzburg model and then describes analytical and numerical results for the bulk phase diagram. Section III presents analytical and numerical results for the surface states, in the absence of a surface field. Section IV shows how the surface phase diagram is modified by a surface field. Finally section V discusses the results and their possible realization. ## II Model and bulk phase diagram In this section we introduce the model used in the rest of the article, and we study the bulk phase diagram, especially the transition line separating the homogeneous and modulated states. The bulk free energy $`F_b`$ is the spatial integral of the density $`_b`$, which is the following functional of the scalar order parameter $`\varphi (x)`$: $$_b[\varphi ]=h\varphi +\frac{1}{2}r\varphi ^2+\frac{1}{4}\varphi ^4\frac{1}{2}(\varphi ^{})^2+\frac{1}{2}(\varphi ^{\prime \prime })^2$$ (1) where $`\varphi ^{}=d\varphi /dx`$. We have scaled the order parameter, the energy and the unit of length to simplify the coefficients, and so $`h`$ and $`r`$ are the rescaled ordering field and temperature variables respectively. The corresponding Euler-Lagrange equation is $$\varphi ^{\prime \prime \prime \prime }+\varphi ^{\prime \prime }h+r\varphi +\varphi ^3=0.$$ (2) Nakanishi and Fisher examined a very different model; the gradient-squared term appeared with a positive coefficient, the $`(\varphi ^{\prime \prime })^2`$ term was omitted, and surface terms were added. Their model, without the surface terms, applies to the usual Ising model with ferromagnetic interactions; it has only the disordered and homogeneous (ferromagnetically ordered) phases, and its bulk $`(r,h)`$ phase diagram consists of a first-order line at $`h=0`$ and $`r<0`$ which terminates at a critical point at $`r=0`$. The model of Equation (1), but without the $`(\varphi ^{})^2`$ term, exhibits a Lifshitz point at $`h=r=0`$ and a first-order line for $`r<0`$. Without the bulk field $`h`$, the model (1) has a disordered phase at high temperature ($`T`$), a second-order transition at $`r=\frac{1}{4}`$ to a modulated phase, and a strong first-order transition at $`r1.2`$ to one of two degenerate homogeneous phases; the modulated phase is almost sinusoidal over its entire range, and its wavenumber is almost independent of $`T`$. In the $`(r,h)`$ plane, the modulated phase occupies a closed loop . Outside this loop, the energetically favored phase is the homogeneous phase, with order parameter $`\varphi _0`$ found from $$h+r\varphi _0+\varphi _{0}^{}{}_{}{}^{3}=0;$$ (3) its free-energy density is $`_0=h\varphi _0+\frac{1}{2}r\varphi _0^2+\frac{1}{4}\varphi _0^4`$. Figure 1 gives the bulk phase diagram, as found for the most part by numerical solution of the Euler-Lagrange equation (2) with periodic boundary conditions. The homogeneous-modulated transition is second-order near $`r=\frac{1}{4}`$, but otherwise first-order. The second-order segment and the tricritical points at its ends are found analytically in the following. We consider a spatially modulated order parameter and expand it in harmonics. If $`q`$ is the wavenumber of the modulated structure and $`ϵ`$ is the amplitude of the leading harmonic, then the order parameter takes the form $$\varphi (x)=\varphi _0+ϵ\mathrm{cos}(qx)+ϵ^2[\varphi _2\mathrm{cos}(2qx)+\overline{\varphi }_2]+O(ϵ^3)$$ (4) where $`\varphi _2`$ and $`\overline{\varphi }_2`$ are constants to be determined. Inserting this order parameter in the free energy (1) and integrating over a period, one finds the following expansion of the free energy (per unit volume): $$_b=_0+ϵ^2_2+ϵ^4_4+O(ϵ^6)$$ (5) with coefficients $$_2=\frac{1}{4}(r+3\varphi _0^2\frac{1}{4}),$$ (6) $`_4`$ $`=`$ $`\frac{1}{4}(r+3\varphi _0^2+2)\varphi _2^2+\frac{1}{2}(r+3\varphi _0^2)\overline{\varphi }_2^2`$ (7) $`+`$ $`\frac{3}{2}\varphi _0\overline{\varphi }_2+\frac{3}{4}\varphi _0\varphi _2+\frac{3}{32}.`$ (8) The free energy has already been minimized with respect to the wavenumber $`q`$, giving $`q=\sqrt{1/2}+O(ϵ^2)`$. The homogeneous phase is unstable to a modulated perturbation when $`_2<0`$. Provided that $`_4>0`$ then, a second-order transition occurs at $$h=\pm \frac{2}{3\sqrt{3}}\left(\frac{1}{8}+r\right)\sqrt{\frac{1}{4}r}.$$ (9) When $`_4`$ is negative, the transition to the modulated phase is first-order. To find the tricritical points separating the continuous and first-order segments, we minimize $`_4`$ with respect to $`\varphi _2`$ and $`\overline{\varphi }_2`$ and then set the result equal to zero. On the line $`_2=0`$, $`_4`$ is minimized by $`\varphi _2=2\varphi _0/3`$ and $`\overline{\varphi }_2=6\varphi _0`$; the minimum value is $$_4=\frac{3}{32}\frac{19}{4}\varphi _0^2,$$ (10) and so the two tricritical points are located at $$r=29/152,h=\pm \sqrt{6/19^3}.$$ (11) ## III Surface phase diagram In this section we consider the surface phase diagram of the model (1) for a semi-infinite system, with no surface field. The presence of the surface generally produces states localized near the surface, and the states are energetically favored in part of the phase diagram. We studied the surface states in the region where the bulk phase is homogeneous, and examined their transitions with varying temperature and the external field. Only a cursory examination was made in the region where the bulk is modulated; in this region, we found many solutions of the Euler-Lagrange equation, so many that a detailed analysis was felt unjustified at this time. That is, surface states and surface phase transitions may exist inside the bulk modulated loop, but have not been studied. We consider a system occupying the half-space $`x0`$, and we assume that the order parameter depends only on $`x`$. The bulk energy $`F_b`$ is found by integrating the density of (1). In this section, we treat the surface very simply, by assuming that it merely terminates the bulk; we thus take free boundary conditions at the surface. In section IV, however, we assume that the surface also applies a local ordering field $`h_s`$; then the total energy is $`F_b+F_s`$, where $$F_s=h_s\varphi _s$$ (12) and $`\varphi _s`$ is the order parameter at $`x=0`$. The general boundary conditions are then $$\varphi _s^{}+\varphi _s^{\prime \prime \prime }h_s=0,\varphi _s^{\prime \prime }=0.$$ (13) We solved the Euler-Lagrange equation (2) numerically subject to the boundary conditions (13). This equation can have many solutions, depending on the bulk field $`h`$ and the temperature variable $`r`$. Figure 2 gives the surface phase diagram for $`h_s=0`$, as found from examining these solutions. The surface states are energetically favorable inside the outer loop of the figure (with the qualification noted above), and the homogeneous states outside; the surface states exist (as solutions of the Euler-Lagrange equation) outside this loop but are only metastable there. An interesting feature is that the surface orders at $`r=1`$ for $`h=0`$, but the bulk orders only at $`r=\frac{1}{4}`$. Many more surface states were found at lower temperatures, but they were always metastable. Figure 3 shows a typical profile of the surface state in the ordered region, for a small and negative bulk field (to break the symmetry) and $`h_s=0`$. The order parameter decays to the bulk value (which is negative) far from the surface, but it is large and positive near it; the overshooting and the damped oscillations result from a complex decay constant, as shown below. Correspondingly, when the bulk field is positive, the order parameter of the surface state is negative near the wall and then decays to the positive bulk value. Thus at $`h=0`$ there is a first-order transition at which the surface state changes sign. To provide an analytical understanding of these numerical results and also those of the next section, we present the following stability analysis of the homogeneous bulk state. The analysis is valid when the deviation of the order parameter from the bulk value is small. The order parameter is written as $`\varphi =\varphi _0+\psi `$, where $`\varphi _0`$ is given by (3) and $`\psi `$ is the deviation. The free-energy density $`=_b_0`$ associated with $`\psi `$ is $$=\frac{1}{2}(r+3\varphi _{0}^{}{}_{}{}^{2})\psi ^2+\varphi _0\psi ^3+\frac{1}{4}\psi ^4\frac{1}{2}(\psi ^{})^2+\frac{1}{2}(\psi ^{\prime \prime })^2.$$ (14) The energy is minimized by an order parameter $`\psi `$ which satisfies the Euler-Lagrange equation $$(r+3\varphi _{0}^{}{}_{}{}^{2})\psi +3\varphi _0\psi ^2+\psi ^3+\psi ^{\prime \prime }+\psi ^{\prime \prime \prime \prime }=0.$$ (15) To prepare for the next section, we include also the surface free energy (12). The boundary conditions are then $$\psi ^{}(0)+\psi ^{\prime \prime \prime }(0)h_s=0,\psi ^{\prime \prime }(0)=0.$$ (16) For $`h_s=0`$, the homogeneous bulk state $`\psi =0`$ is clearly a solution of Equations (15) and (16). This solution is stable over some region of the $`(r,h)`$ plane, but it becomes unstable at the transition to the bulk modulated state. To study the surface states, we solve Equations (15) and (12) perturbatively in $`\psi `$. The expansion starts from the solution $$\psi _1(x)=Ae^{\alpha x}+A^{}e^{\alpha ^{}x}$$ (17) of the linearized Equation (15). The amplitude $`A`$ and the decay constant $`\alpha `$ are both complex; the latter (with positive real part) is found from $$\alpha ^2=\frac{1}{2}(1+i\gamma )$$ (18) where $`i=\sqrt{1}`$ and $`\gamma =[4(r+3\varphi _0^2)1]^{1/2}`$. The condition $`\psi _1^{\prime \prime }(0)=0`$ in (16) gives the amplitude $`A`$ in terms of $`m=\psi (0)`$ as $$A=\frac{m}{2}\left(1\frac{i}{\gamma }\right).$$ (19) It is convenient to take $`m`$ as the expansion parameter. The solution (17) gives the free energy to order $`m^2`$. In order to obtain the free energy to the required order ($`m^4`$), one must find the higher-order contributions to $`\psi `$. Let $`\psi =\psi _1+\psi _2`$, where $`\psi _2`$ is the nonlinear part of $`\psi `$. Inserting this form in Equation (15), using (17) and keeping terms to $`m^3`$, one finds $`\psi _2(x)`$ $`=`$ $`B_1e^{2\alpha x}+\frac{1}{2}B_2e^{(\alpha +\alpha ^{})x}+C_1e^{3\alpha x}`$ (20) $`+`$ $`C_2e^{(2\alpha +\alpha ^{})x}+De^{\alpha x}+c.c.`$ (21) with $`\psi _2(0)=\psi _2^{\prime \prime }(0)=0`$ and coefficients $`B_1`$ $`=`$ $`\left(3+5\sqrt{3}i\right)\varphi _0m^2/126,`$ (22) $`B_2`$ $`=`$ $`2\varphi _0m^2/3,`$ (23) $`C_1`$ $`=`$ $`\left(27+11\sqrt{3}i\right)m^3/13104,`$ (24) $`C_2`$ $`=`$ $`\left(3+2\sqrt{3}i\right)m^3/84,`$ (25) $`D`$ $`=`$ $`\left({\displaystyle \frac{5}{14}}{\displaystyle \frac{19}{126}}\sqrt{3}i\right)\varphi _0m^2\left({\displaystyle \frac{7}{208}}+{\displaystyle \frac{47}{4368}}\sqrt{3}i\right)m^3.`$ (26) On using the result $`\psi =\psi _1+\psi _2`$ in the free energy (14) and integrating over $`x`$, one finds that the free energy of the surface state (per unit area) is given by $$F=h_sm+a_2m^2+\frac{2}{9}\varphi _0m^3+\frac{3}{56}m^4+O(m^5)$$ (27) where $$a_2=\left(r+3\varphi _0^2\right)\frac{i}{2\gamma }\frac{\alpha ^3\alpha ^3}{\alpha \alpha ^{}}.$$ (28) The amplitude of the surface structure is determined by minimizing the free energy with respect to $`m`$ for given surface field $`h_s`$. This amounts to satisfying the first condition in Equation (16). We now use the free energy (27) to discuss the surface phase diagram in the region where $`m`$ is small. Consider first the case $`h_s=0`$. For zero bulk field $`h`$, $`\varphi _0=0`$ and there is no surface state when $`a_2>0`$ (that is, $`\psi =0`$). Setting $`a_2=0`$, one finds a continuous transition at $`r=1`$ from the disordered bulk state $`m=0`$ to a surface state with $`m0`$; this is the second-order point at the right of Figure 2. For field $`h0`$, the bulk order parameter $`\varphi _0`$ is also non-zero and the free-energy expansion (27) has a term in $`m^3`$; this cubic term gives a first-order transition to the surface state, again as found numerically. Near the point $`(r=1,h=0)`$, $`m`$ is small and the transition line can be found approximately from the free-energy expansion (27). Away from this point, however, the full free energy must be minimized numerically; Figure 2 gives the resulting $`(r,h)`$ surface phase diagram for $`h_s=0`$. ## IV Effect of a surface field We consider now the surface phase diagram for non-zero surface field $`h_s`$. Positive $`h_s`$, for example, tends to increase the order parameters of all states in the region near the surface. The new feature is that the Euler-Lagrange equation must now be solved numerically for what were homogeneous bulk states at $`h_s=0`$; for lack of a better term, we refer to these surface-field-modified bulk states simply as bulk states. Figures 4 and 5 give parts of typical phase diagrams for $`h_s>0`$, as found by numerical solution of Equation (2), subject to the boundary conditions (13). Figure 4 shows the high-temperature part of the phase diagram for $`h_s=10^4`$. The surface field breaks the transitions of Figure 2 into two first-order lines at which the surface state changes discontinuously. In the region bounded by the upper line and the left vertical (where the bulk field $`h`$ is positive), the order parameter of the surface state is negative at the boundary $`x=0`$ ($`\varphi _s<0`$). The lower line ends at a second-order point. Below this point there is a first-order transition between the paramagnetic state and the surface state with $`\varphi _s>0`$, while above it the two states are indistinguishable. The free-energy expansion (27) can be used to find this point to leading order in $`h_s`$; the result is $`a_2`$ $`=`$ $`{\displaystyle \frac{3}{2}}\left({\displaystyle \frac{3}{14}}\right)^{1/3}h_s^{2/3},`$ (29) $`\varphi _0`$ $`=`$ $`{\displaystyle \frac{9}{2}}\left({\displaystyle \frac{3}{14}}\right)^{2/3}h_s^{1/3},`$ (30) in good agreement with the numerical results. Figure 5 shows the low-temperature part of the phase diagram for $`h_s=10^1`$. Paradoxically, a positive surface field cooperates, rather than competes, with a negative bulk field to enhance the stability of the lower surface state (and it competes with a positive bulk field for the other). These effects occur because the order parameter of the surface state changes sign (as seen in Figure 3). ## V Discussion We have developed and analyzed a model to describe the effect of a substrate (or a free surface) on a material which has a bulk phase transition between homogeneous and modulated states. Modulated states tend to form because the free energy of the model contains a term, quadratic in first derivatives of the order parameter, which has a negative coefficient. We treated the surface first as simply terminating the bulk, and then in addition as supplying a surface field coupling linearly to the order parameter. The important new result of our analysis is the quite unexpected existence of solutions localized at the surface, solutions which exist even if the surface field is zero. These solutions are energetically favored for temperature and field values that are outside but not too far from the closed loop within which the modulated bulk state is stable. When the surface field differs from zero, the loop breaks apart (as shown in Figures 4 and 5). We now turn our attention to the applicability of our results to cholesteric liquid crystals in a field . It is obviously desirable to estimate the conditions of chirality, temperature, field, and surface interactions for which the surface states should be observable. To do this, we should examine the relationship between the variables of the theory and the experimental variables, by comparing the expressions for $`_2`$ and $`_4`$ in section II of this paper with the analogous expressions in section III-C of . It is reasonable however, and far simpler, to expect the loop regions to scale by the same factors; this should be true independent of the strength of surface interactions. From Figures 1 and 2, the outer (surface-state) loop extends over the range $`4.5r<1`$ while the inner loop extends over $`1.2r<\frac{1}{4}`$, about a factor of four. Accordingly, we estimate the surface-state region to be four times the size of the undulating-state region in temperature. From , the undulating state should occur for intrinsic pitches in the range of $`1260630`$ nm, at electric fields of the order of a few hundreds of kV/cm, or magnetic fields of roughly $`40`$ T; the temperature width was estimated to be a few tenths of a degree. These conditions are very difficult to achieve and account for the fact that the undulating state has not yet been observed, although some groups plan to attempt the experiments. The surface-state region is expected to be one degree wide. Techniques sensitive to birefringence near the surface, such as Brewster-angle ellipsometry , may be able to detect the surface states. The surface states should appear in incommensurate systems where the modulated phase is driven by a negative gradient-squared term. Conditions may be favorable in magnetic Lifshitz-point materials like MnP, or in Langmuir monolayers or diblock copolymers. Other systems in which a modulated phase is driven by a negative gradient-squared term are sodium nitrite and thiourea ; related systems are quartz and berlinite, but for these the modulated phase is two-dimensional. ###### Acknowledgements. We thank R. Seidin for helpful discussion of the bulk phase diagram and R. C. Desai for helpful comments. This research was supported by the National Science Foundation under Science and Technology Center ALCOM Grant No. DMR 89-20147, the Einstein Center for Theoretical Physics, the Inter-University High Performance Computation Center (Tel Aviv), the Natural Sciences and Engineering Research Council of Canada, and the Meyerhoff Foundation.
no-problem/9908/astro-ph9908333.html
ar5iv
text
# 1 Introduction ## 1 Introduction Calibrations of the form $`a_P\mathrm{log}P+a_C\mathrm{color}+a_0=M_V`$ are relatively successful in predicting absolute magnitudes of contact binaries of the W UMa-type. They have been established on the basis of systems in open clusters (Rucinski 1994a) and then improved for the solar neighborhood using the Hipparcos data (Rucinski and Duerbeck 1997); with metallicity corrections, they can be used for globular clusters as well (Rucinski 1994b, 1995). They have been invaluable in handling the Galactic Disk and open cluster data for contact binaries (Rucinski 1998a). It would be very useful to establish a similar calibration for contact binaries of early spectral types because such binaries are now being detected in large numbers in nearby galaxies (Kałużny et al. 1998, Stanek et al. 1998, Stanek et al. 1999) and may eventually provide independent distance estimates. Contact binaries of spectral earlier than middle-A and periods longer than about 1.5 day are very rare in the Galaxy. They practically do not exist in the OGLE Galactic-Bulge database which probes mostly the Old Disk population (Rucinski 1997a, 1998a); the period distribution of contact binaries abruptly ends at about 1.3–1.5 days (Rucinski 1998b). This cut-off can be explained by non-existence of Old Disk main-sequence stars with masses much larger than the solar mass. Intrinsically bright, massive contact binaries, such as the systems discussed by Popper (1982) – with periods of a few days and masses of several solar masses – do exist within our Galaxy, but their frequency of occurrence is apparently exceedingly low and currently unknown. It is important that they share the same property of identical effective temperatures of components – in spite of strongly differing masses – with the solar-type contact binaries of the W UMa-type. This property basically defines the W UMa-type stars. The currently ongoing microlensing projects lead to discoveries of large numbers eclipsing binaries in Magellanic Clouds. This paper presents a pilot study based on the data for contact binaries in the Large Magellanic Cloud (LMC) which have been collected by the MACHO project (Alcock et al. 1997). In the next sections, we describe various properties of the early-type contact systems in the MACHO sample. The electronically-available database (American Astronomical Society CD-ROM Series, Vol. 8, 1997) contains the summary, time-independent data and plots of the light curves, and is quite typical for microlensing projects. It can be treated similarly as the one obtained by the OGLE project for the direction toward the Galactic Bulge (Rucinski 1997a). In this spirit, only the mean quantities such as the period, the maximum and the in-eclipse $`V`$-magnitudes as well as the maximum-light $`VR_C`$ color index in the Kron–Cousins system have been used. Availability of this particular index is somewhat unfortunate because the previous absolute-magnitude calibrations for W UMa-type systems were based on the $`BV`$ and $`VI_C`$ indices so that the existing calibrations can be used only for general guidance, but not for detailed comparisons. ## 2 The Sample The definition of W UMa-type contact systems (normally abbreviated as EW) does not include binaries with periods longer than one day. In the MACHO database of the eclipsing binaries discovered in LMC (Alcock et al. 1997) such systems are called EB, with a subset of them, with eclipses of similar depth indicating similar surface brightness of components, designated as EB3. The EB3 systems have been used in this paper, keeping in mind subjectivity of this definition. Electronically available plots of the light curves confirm, that the light-curve shapes are indeed very typical for contact binaries, as shown for the example in Fig. 7 in Alcock et al. (1997). There are 86 systems of this type in the MACHO database. In addition to the values of the orbital periods, the photometric data are important for discussions of this paper. The precision of the MACHO data is moderate: The mean standard error of $`V`$ magnitudes is about 0.07 while the mean standard error for the color index $`VR_C`$ is about 0.03. The accuracy (which reflects the role of systematic errors) is limited for the MACHO sample by the use of the non-standard bandpasses which are not defined – as usually – by glass filters, but by a dichroic filter separating the bands, and by transmission characteristics of the CCD’s and of the atmosphere. The histograms of the $`V`$-magnitude amplitudes and minimum differences shown in Fig. 1 are quite typical for contact systems. The largest amplitudes observed are slightly over 0.8 mag and the minimum differences are typically in the range of 0 to 0.02 mag, as for genuine W UMa-type systems (see Fig. 2 in Rucinski 1997b; the magnitude difference is approximately twice the Fourier coefficient $`a_1`$ shown there). The limit for inclusion in the database was the variability amplitude larger than 0.2 mag. Contact binaries with still smaller variability amplitudes are expected to be actually the most frequent for randomly distributed orbital inclinations (Rucinski 1997b). ## 3 The Color–Magnitude Diagram (CMD) The CMD for the contact systems considered here shows a concentration of the systems at the left, blue edge of the diagram and a scattered population of systems in the red part of the diagram (Fig. 2). Since the available data are in two photometric bands only, we have no information about reddening of individual objects. Reddening in LMC is patchy and varies within $`0<E_{BV}<0.4`$, but with excursions up to 0.8 (Harris et al. 1997). In this situation, it has been decided to limit our considerations to the systems at the left edge of the CMD which are intrinsically the bluest and least-reddened. The systems in the red part of the diagram in their majority have long orbital periods (see the next Section and Fig. 3) and we do not have a good explanation for them. In what follows, we will consider three groups of the contact binaries defined by the observational $`VR_C`$ color ranges: the extremely blue (XB) systems with $`VR_C<0.1`$, the moderately blue (MB) systems with $`0.1<VR_C<0`$ and the red systems with $`VR_C>0`$; the latter are not discussed here at all. ## 4 The Period–Color Diagram (PCD) The period-color diagram is shown in Fig. 3, with the same color-range symbols as in Fig. 2. The curved line at the left edge gives the Short-Period Blue-Envelope (SPBE) for the Galactic-Disk sample, $`(VI_C)=0.053P^{2.1}`$ (Rucinski 1997a), after application of the color-color transformation to $`VR_C`$ using Table 3 in Taylor (1986). It is obvious that the LMC systems are seen beyond the blue limit of the SPBE which was previously found to obey for all contact binaries in the galactic field and in all open clusters. The low metallicity of LMC ($`[\mathrm{Fe}/\mathrm{H}]=0.5`$) cannot produce the shift because the intrinsic color indices for main-sequence OB stars have a negligible dependence on metallicity (Oestreicher et al. 1995). We strongly suspect that by considering the bluest contact systems, we are also looking at objects of a young stellar population which simply does not exist in the solar neighborhood or in old open galactic clusters. It is possible, that the curvature of the SPBE that was found before was entirely due to the lack of young objects in the previously-used samples; possibly, the real location of the SPBE is closer to what has been marked by a broken line in Fig. 3. A closer inspection of the PCD in Fig. 3 shows that the blue contact systems in LMC have orbital periods up to 2–3 days, whereas in the disk sample of OGLE the period distribution was found to a sharp cut-off at about 1.3–1.5 days (Rucinski 1998b). There are too few systems in the MACHO sample to analyze statistically the numbers of systems with periods above 1.3 days, however. ## 5 The Period–Luminosity Relation The contact systems in LMC are practically in the same distance from us so that the initial goal of this study was an attempt to establish an absolute-magnitude calibration of the type $`a_P\mathrm{log}P+a_{VR}(VR_C)+a_0=M_V`$. However, this goal is impossible to achieve at this time for the following reasons: (1) Lack of reddening information prevents determination of reddening and absorption corrections for individual systems; (2) $`VR_C`$ color index, as other ones utilizing optical spectral bands, loses sensitivity to the effective temperature for very hot stars and hence must be determined with high accuracy; (3) The observed color indices have modest precision of about 0.03 mag and an uncharacterized accuracy due to the use of the non-standard bandpasses. At present, one can only address the matter of period dependence in the observed values of $`V_{\mathrm{max}}`$. Such a limited goal is still a valuable check on the assumptions because a period dependence is expected only for genuinely contact systems and it is by no means obvious that the systems classified as EB3 are indeed contact ones. For detached binaries, no relation between period and component brightness is expected. The period – observed-magnitude diagram for the sample is shown in Fig. 4. As one can see, the systems of both groups, the extra-blue (XB) and the moderately-blue (MB) systems, show some period dependence, but the scatter of points is large, probably mostly because of the unaccounted effects of the interstellar reddening and extinction. We note also that there are more MB systems in the vicinity of the faint-limit of the sample at $`V_{\mathrm{max}}18`$ than XB systems. Least-squares fits of the form $`V_{\mathrm{max}}=a_0+a_1\mathrm{log}P`$ have been performed for both samples; we note that the orbital periods $`P`$ are known practically without errors. Because the scatter of the data is large, uncertainties of the coefficients crucially depend on the data sampling. To characterize this effect, the errors of the coefficients $`a_i`$ have been determined using the bootstrap re-sampling technique. The results are listed in Table 1 in terms of the median values and the 68 percent (for Gaussian distributions, $`\pm 1`$-sigma) and 95 percent ($`\pm 2`$-sigma) confidence levels. Histograms of the individual values of the zero-point and slope coefficients from the bootstrap experiment are shown in Fig. 5. Note that the mean color indices for the XB and MB groups are $`\overline{VR_C}=0.142`$ and $`0.047`$, while the difference of the zero points is: $`a_0(\mathrm{MB})a_0(\mathrm{XM})=0.84`$. This would imply a very strong dependence on the color index with a steep slope of about 8.8, but part of this is almost certainly contributed by the relatively larger reddening for the MB group. Much more interesting are the period-dependence slope coefficients $`a_1`$. The determination of $`a_1`$ for the XB group is surprisingly stable, given the large scatter of the data points and the fact that the value of $`a_1`$ is driven mostly by the outlying points, $`a_1(\mathrm{XB})=2.74_{0.68}^{+0.58}`$; the range given here is equivalent to the rms error of 1-sigma. The determination for the MB group is poorer, $`a_1(\mathrm{MB})=2.07_{1.00}^{+0.83}`$, probably because the presence of the faint cut-off in the data and certainly stronger influence of the scatter in the reddening values for this group. When compared with the calibrations for the W UMa-type binaries using the $`BV`$ and $`VI_C`$ indices (cf. Eqs. (2) and (5a) in Rucinski and Duerbeck 1997), the slope $`a_1(\mathrm{XB})=2.74`$ is rather shallow (the previous calibrations implied the slope of about $`4.4`$) which may have resulted from relatively large photometric errors of the MACHO data. ## 6 Conclusions for the Future The main result of this pilot study is that the sample of EB3 binaries discovered and classified by the MACHO project certainly contains hot, blue, massive binaries of the contact type. Although it may have been preferable to make the selection of the systems on the basis of the Fourier coefficients – rather than to rely on the MACHO classifications – this was not necessary because the systems reveal typical properties of contact-binary stars. The contact nature of these binaries is most strongly confirmed by the existence of a period–luminosity relation which is best visible in the XB sub-sample consisting of the bluest and least reddened systems. The new result is that the massive, young, blue systems in LMC are apparently not constrained by the previously-established short-period blue envelope in the period–color diagram and appear with blue-color/short-period combinations not observed in the previous surveys of the open clusters in the Galaxy and in the Galactic Disk field. These systems appear also within the orbital period interval of 1.3–1.5 to 2–3 days, a range for which the frequency of the galactic contact systems is known to be un-measurably low. An important conclusion related to the future attempts of determining an absolute-magnitude calibration for early-type contact systems is the availability of accurate color indices. Photometry should be available in at least three bandpasses for reddening determinations, and must be accurate enough (through the use of the standard filter bands and a requirement that mean standard errors be $`<0.01`$ mag) to compensate for the decreased sensitivity of optical color indices to effective temperature. For comparison and consistency checks with the results for solar-type contact binaries of the W UMa-type, it would be preferable to utilize $`BV`$ and $`VI_C`$ color indices. The new database for SMC coming from the OGLE-II project (Udalski et al. 1998) apparently fulfills most of the above desiderata. Acknowledgements. Thanks are due to Bohdan Paczynski for pointing the existence of the MACHO data for eclipsing binaries in LMC and for very useful comments and suggestions on the first version of the paper. ## REFERENCES * Alcock, C., et al. 1997, Astron. J., 114, 326. * Harris, J., Zaritsky, D., and Thompson, I. 1997, Astron. J., 114, 1933. * Kałużny, J., Stanek, K.Z., Krockenberger, M., Sasselov, D.D., Tonry, J.L., and Mateo, M. 1998, Astrophys. J., 115, 1016. * Oestreicher, M.O., Gochermann, J., and Schmidt-Kaler, T. 1995, Astron. Astrophys. Suppl. Ser., 112, 495. * Popper, D.M. 1982, Astrophys. J., 262, 641. * Rucinski, S.M. 1994a, P.A.S.P., 106, 462. * Rucinski, S.M. 1994b, Astron. J., 107, 738. * Rucinski, S.M. 1995, P.A.S.P., 107, 648. * Rucinski, S.M. 1997a, Astron. J., 113, 407. * Rucinski, S.M. 1997b, Astron. J., 113, 1112. * Rucinski, S.M. 1998a, Astron. J., 116, 2998. * Rucinski, S.M. 1998b, Astron. J., 115, 1135. * Rucinski, S.M., and Duerbeck, H.W. 1997, P.A.S.P., 109, 1340. * Schmidt-Kaler, T. 1982, in: Landolt-Börnstein, New Series, Vol. 2b, Astron. Astrophys., Ed. Springer-Verlag, p.21. * Stanek, K.Z., Kałużny, J., Krockenberger, M., Sasselov, D.D., Tonry, J.L., and Mateo, M. 1998, Astrophys. J., 115, 1894. * Stanek, K.Z., Kałużny, J., Krockenberger, M., Sasselov, D.D., Tonry, J.L., and Mateo, M. 1999, Astrophys. J., 117, 2810. * Taylor, B.J. 1986, Astrophys. J. Suppl. Ser., 60, 577. * Udalski, A., Soszyński, I., Szymański, M., Kubiak, M., Pietrzyński, G., Woźniak, P., and Żebruń, K. 1998, Acta Astron., 48, 563.
no-problem/9908/cond-mat9908371.html
ar5iv
text
# Percolative phase separation induced by nonuniformly distributed excess oxygens ## Abstract The zero-field <sup>139</sup>La and <sup>55</sup>Mn nuclear magnetic resonances were studied in $`\mathrm{La}_{0.8}\mathrm{Ca}_{0.2}\mathrm{MnO}_{3+\delta }`$ with different oxygen stoichiometry $`\delta `$. The signal intensity, peak frequency and line broadening of the <sup>139</sup>La NMR spectrum show that excess oxygens have a tendency to concentrate and establish local ferromagnetic ordering around themselves. These connect the previously existed ferromagnetic clusters embedded in the antiferromagnetic host, resulting in percolative conduction paths. This phase separation is not a charge segregation type, but a electroneutral type. The magnetoresistance peak at the temperature where percolative paths start to form provides a direct evidence that phase separation is one source of colossal magnetoresistance effect. Since the discovery of colossal magnetoresistance (CMR) effect in $`\mathrm{La}_{1\mathrm{x}}\mathrm{Ca}_\mathrm{x}\mathrm{MnO}_3`$ (LCMO), many theoretical and experimental works have been done to find the physical mechanism of CMR effect because of their interesting physical properties and application potential. The first explanation of the most interesting physical property of LCMO, the simultaneous occurrence of the paramagnetic to ferromagnetic and insulator to metal transitions, was the simple double exchange model given by Zener in 1951. However, Millis pointed out that the resistivity of Sr-doped manganites cannot be fully explained by double exchange alone in 1995. Thereafter, several theories have been proposed to describe the physical properties of CMR materials more completely, one of which is phase separations(PS). In the low doping range ($`x0.2`$), the magnetic phase of LCMO is not homogeneous. The existence of magnetic PS was verified by the simultaneous observation of ferromagnetic and antiferromagnetic nuclear magnetic resonance (NMR) signals at low temperature. Since the ferromagnetic metallic regions are embedded in the antiferromagnetic insulating regions, LCMO in this range shows ferromagnetic insulating behavior macroscopically. Since then, PS has been suspected as one of the possible mechanisms of CMR effect. On the other hand, PS has been also observed near phase transition temperature in LCMO for $`0.2<x<0.5`$, which are homogeneous ferromagnetic metals well below phase transition temperature. These two kinds of PS are thought to be originated from different mechanisms, because the PS observed in the low doping range is the ground state and stable in a wide temperature range, while the PS observed for $`0.2<x<0.4`$ occurs only near phase transition temperature. In this report, we will focus our discussion on the PS in low doped LCMO. Thoery predicts two different types of PS in low doped LCMO. One is the charge segregation type and the other is the electroneutral type. Yunoki studied the 2-orbital Kondo model including the classical Jahn-Teller phonons and found that PS is induced by the orbital degrees of freedom. In such a case, the charge density of $`\mathrm{e}_\mathrm{g}`$ electron is not stable at a special value of chemical potential, resulting in two regions with different charge densities. The size of both regions is expected to be very small, about the order of nanometer, due to the extended Coulomb interaction. Recently, Uehara presented TEM images of $`\mathrm{La}_{5/8\mathrm{y}}\mathrm{Pr}_\mathrm{y}\mathrm{Ca}_{3/8}\mathrm{MnO}_3`$ which shows the mixture of the charge ordering phase of $`\mathrm{La}_{0.5}\mathrm{Ca}_{0.5}\mathrm{MnO}_3`$ type and ferromagnetic phase at low temperature. The sizes of both regions are about 0.5 $`\mu m`$, which is too large to be explained by the charge segregation type PS. On the other hand, Nagaev paid attention to the PS induced by nonuniformly distributed oxygens. He pointed out that the regions enriched with oxygen have an enhanced hole density and the holes establish local ferromagnetic ordering, whereas the remained regions are poorly conductive and antiferromagnetic due to the electron-hole recombination. In this case, PS is the electroneutal type because the densities of holes and excess oxygens are same in a given region, and the region sizes can be much larger than those of the charge segregation type PS. Several reports have supported the existence of PS, but it still remains unclear which scenario is more correct in low doped LCMO. There have been many works which showed that oxygen plays an important role in determining electromagnetic properties of LCMO, but it’s effects on PS have never been studied. In this work, we report that the electroneutral percolative PS is formed in low doped LCMO by excess oxygens. Experimental results provide an evidence for the fact that the PS in the low doping range is one source of CMR. Two polycrystalline samples of $`\mathrm{La}_{0.8}\mathrm{Ca}_{0.2}\mathrm{MnO}_{3+\delta }`$ with different $`\delta `$ values were synthesized by the conventional solid state reaction method. The starting materials were La<sub>2</sub>O<sub>3</sub>, MnCO<sub>3</sub>, and CaCO<sub>3</sub>. Calcining and sintering with intermediate regrinding were repeated in the temperature range of 1000 <sup>0</sup>C - 1350 <sup>0</sup>C for four days. Sample 1 was obtained by annealing in air at 1100 <sup>0</sup>C for two days, and sample 2 was obtained by additional grinding, sintering and annealing of a part of sample 1 in oxygen flow (200 cc/min) at the same temperatures with sample 1. The crystal structures were examined by x-ray powder diffraction with Cu $`K\alpha `$ radiation. Both samples were single phase and orthorhombic. The lattice parameters of sample 1 and 2 were $`a_1`$=5.489 Å, $`b_1`$=5.496 Å, $`c_1`$=7.765 Å, and $`a_2`$=5.502 Å, $`b_2`$=5.507 Å, $`c_2`$=7.777 Å, respectively. Resistivity was measured using the conventional four-probe method, and magnetization was measured by a commercial SQUID magnetometer. Zero field NMR spectra were obtained by using a spin-echo technique. Fig. 1 shows the temperature dependence of magnetization at 1 Tesla. The paramagnetic Curie temperatures (T<sub>C</sub>) of sample 1 and 2 are 191.5 K and 192.2 K, respectively. The difference of T<sub>C</sub> values less than 1 K indicates that the difference between $`\delta `$ values of sample 1 and 2 is less than 0.02. The magnetizations of sample 1 and 2 at low temperature were almost same, and the magnetic field dependences of the magnetizations were almost same either. Though the macroscopic magnetic properties of two samples are very similar quantitatively, local magnetic environments are quite different as seen in the <sup>139</sup>La NMR spectra obtained at 78 K (Fig. 2). In the figure, two differences are noticeable between the spectra of sample 1 and 2. First, the signal intensity of sample 2 is about five times that of sample 1, and second, the resonance frequency of sample 2 is higher and the linewidth is much broader, especially in the high frequency side. We discuss about the difference of signal intensity first. The NMR signal intensity of a ferromagnet in zero field is proportional to $`\eta VH_L/T`$, where $`\eta `$ is the enhancement factor, V is the volume of the ferromagnetic region of a sample, and $`H_L`$ is the local field at the nuclei of interest. Since magnetization which is proportional to V does not change, the NMR signal intensity change at a given temperature and a frequency by extra oxygens is due to the change of the enhancement factor. Fig. 3 provides experimental evidences for this claim. Fig. 3(a) shows that sample 2 gives the maximal signal at the much lower rf power than sample 1, meaning that rf field is more enhanced in sample 2 than in sample 1. The rf enhancement factors of sample 1 and 2 are about 22 and 105, respectively. Fig. 3(b) shows the normalized NMR signal intensity vs. external magnetic field obtained at the fixed rf power which makes the maximal signal in zero field. In this figure, we notice that the signal of sample 2 decays almost to zero while that of sample 1 decays slowly approaching the saturation field, about 3 koe. These are the typical responses of single and multi-domain ferromagnets, respectively. The signal of sample 1 decays a little because the enhancement factor decreases with external field. On the other hand, the drastic signal decay of sample 2 means that domain walls disappear approaching the saturation field. Therefore, these results show that sample 2 has domain walls while sample 1 does not. The NMR signal of sample 2 comes mostly from domain walls because the enhancement factor is usually orders of magnitude larger in domain walls than in domain in general. The size of ferromagnetic clusters embedded in the antiferromagnetic host of sample 1 is not large enough to form multi-domain state. The local ferromagnetic orderings generated by excess oxygens connect some of these ferromagnetic clusters and domain walls are formed on them. In fact, these connected ferromagnetic clusters make also percolative conduction paths as shown in Fig. 4 displaying the temperature dependence of resistivity. The resistivity of sample 1 shows an insulating behavior except a small bending near phase transition, while that of sample 2 shows a broad peak in the temperature range of 170 K - 140 K and a metallic behavior below 140 K. The metallic behavior of sample 2 at low temperature implies that electric transport paths are formed by excess oxygens. The temperature dependence of the La NMR signal intensity of sample 2 shown in Fig. 5 supports the simultaneous generation of conduction paths and domain walls. The signal intensity of homogeneous ferromagnets such as LCMO for $`0.2<x<0.5`$ well follow Curie’s $`T^1`$ law except in the narrow region near T<sub>C</sub>. However, the signal intensity of sample 2 decreases much faster than T<sup>-1</sup>, and almost disappears near 140 K where the metallic behavior fades out (Fig. 4). This means that the total volume of domain walls decreases as temperature increases and the ferromagnetic and metallic conduction paths vanish near 140 K. As approaching this temperature from below, ferromagnetic clusters are disconnected and therefore conduction paths are broken continuously. While the magnetoresistance (MR) curve of stoichiometric perovskite manganite crystals show only one peak near T<sub>C</sub>, that of sample 2 shows one more peak near 140 K as seen in Fig. 4. The MR peak near 170 K is an ordinary CMR peak due to the suppression of spin fluctuation by external field, while the peak near 140 K is undoubtedly related with PS. It is worthwhile to note that the MR near the temperature where the percolative PS is induced is as large as that near the phase transition temperature. External field helps connecting clusters somehow. We now discuss the second difference of the spectra of sample 1 and 2, the difference in resonance frequency and linewidth. The local field $`H_L`$ at the position of a non-magnetic La<sup>3+</sup> ion can be described as $$H_L=A\underset{j}{}n_j\mu _j+H_{dd},$$ where $`A`$ is the transferred hyperfine coupling constant and $`n_j`$ is the number of the j-site Mn moments $`\mu _j`$, surrounding the La ion. $`H_{dd}`$ is the dipolar field summed over all Mn magnetic moments. In perovskite manganites, the dipolar field is negligible and the main contribution to $`H_L`$ comes from the transferred hyperfine field. The transferred hyperfine field is thought to be produced by the $`\pi `$ type overlapping between the Mn $`t_{2g}`$ electron wave function and the oxygen $`|2p_\pi `$ wave function, and the $`\sigma `$ bonding of the oxygen with the $`|sp^3`$ hybrid states of the La<sup>3+</sup> ion. That is, an indirect transferred hyperfine field of the Fermi contact type is mediated by oxygens. Therefore, the constant $`A`$ is a function of the distance between oxygens and a La<sup>3+</sup> ion, and the number of the oxygens surrounding the La<sup>3+</sup> ion. $`n_j`$ and $`\mu _j`$ are almost same in two samples. The distance between oxygens and a La<sup>3+</sup> ion is not an important factor making the difference of the NMR peak frequencies because the peak frequency of sample 2 is higher than that of sample 1 even though the lattice constants of sample 2 are slightly larger than those of sample 1. Therefore, the difference of peak frequencies should be attributed to the difference of the number of oxygens surrounding a La<sup>3+</sup> ion. That is to say, the peak frequency of sample 2 is increased due to the excess interstitial oxygens. There is a report that LaMnO<sub>3+δ</sub> with excess oxygens is characterized by cation vacancy in La and Mn sites rather than by interstitial anions. In this case, however, the lattice parameters decrease as $`\delta `$ increases contrary to our case. Moreover, the NMR spectrum of LaMnO<sub>3+δ</sub> having cation vacancy is well fitted by a single gaussian curve while our sample 2 is not as discussed below. The La NMR spectrum of sample 2 is asymmetric and broader than that of sample 1 in the high frequency side. Since $`\delta `$ is less than 0.02, homogeneous distribution of oxygens cannot enhance the signal in the high frequency side as much as the spectrum in Fig. 2. Therefore, oxygens aggregate to make local ferromagnetic orderings in consistence with Nagaev’s claim that oxygens have a tendency to concentrate. One of the reasons why conduction paths are continuously disconnected as approaching 140 K could be the distribution of excess oxygens becoming more and more uniform as temperature increases. Contrary to La nuclei, the local field at Mn nuclei is negligibly influenced by the local distribution of oxygens because the hyperfine field of the direct Fermi contact type generated by it’s own 3$`d`$ electrons is much stronger. Therefore, the gaussian shape of the Mn NMR spectrum does not change by the presence of excess oxygens as shown in the inset of Fig. 2. The <sup>55</sup>Mn NMR spectra are motionally narrowed by the fast hopping of e<sub>g</sub> electrons between Mn<sup>3+</sup> and Mn<sup>4+</sup> ion sites. Only the Mn nuclei in ferromagnetic regions with the delocalized e<sub>g</sub> electrons contribute to the <sup>55</sup>Mn NMR signal. The local field at Mn nuclei is proportional to the number of average delocalized e<sub>g</sub> electrons. The peak frequency shift by excess oxygens is not less than 10 Mhz. If the distribution of holes is uniform, such a shift corresponds to $`\delta `$0.075, while $`\delta `$ of sample 2 is less than 0.02. This means that the holes are concentrated in ferromagnetic regions. The results of La and Mn NMR imply the concentrated oxygens and holes in ferromagnetic regions, respectively. This support the fact that the PS in low doped LCMO is the electroneutal type. Moreover, considering the easy formation of conduction paths by the aggregation of excess oxygens less than 0.7 %, the size of ferromagnetic clusters are not so small as predicted by the charge segregation type PS. In conclusion, the excess oxygens have a tendency to aggregate and change surroundings into ferromagnetic phase. These local ferromagnetic regions connect ferromagnetic clusters previously existed in stoichiometric samples, which are suspected to be also generated by inhomogeneous distribution of oxygens. This connection produces percolative conduction paths on which domain walls are formed. The observed PS is rather a electroneutral type than a charge segregation type. As temperature increases, the MR peak was observed at the temperature where the percolative PS disappears in addition to the ordinary peak near the phase transition temperature.
no-problem/9908/cond-mat9908029.html
ar5iv
text
# Skyrmion Dynamics and NMR Line Shapes in QHE Ferromagnets ## I INTRODUCTION In the presence of a strong magnetic field, a two-dimensional interacting electron gas (2DEG) exhibits many different quantum states depending on field strength, electron density, and disorder. At Landau level filling factor $`\nu =1`$, when the number of electrons is equal to the number of available Landau orbitals in the lowest Landau level, the 2DEG is in an itinerant ferromagnetic state where all the electron spins are aligned with the magnetic field. The novel features of this quantum Hall ferromagnet (QHF) state originate from the relative strength of the electron-electron Coulomb energy, $`e^2/ϵl_B`$, and the Zeeman energy, $`g\mu _BB`$, which are of order 160 K and 3 K respectively in typical experimental situations. The Pauli exclusion principle combined with the overwhelming cost of Coulomb excitations makes the single-particle spin-$`1/2`$ electron excitation gapped at filling factor $`\nu =1`$. The low energy (but still gapped) charge excitations of the system are Skyrmion spin textures containing many (4-30) flipped spins and are topologically stable. By paying the lower Zeeman price the spins can align locally creating a more advantageous charge distribution that lowers both the Hartree and exchange Coulomb cost with respect to a single spin flip excitation. Because these excitations are the cheapest way to introduce charge into the system, they are present in the ground state of the system at filling factors close to $`\nu =1`$. This produces a rapid reduction of the electron spin polarization away from $`\nu =1`$ which is observed in experiments. The probe of choice to study the electron spin polarization of the 2DEG is nuclear magnetic resonance (NMR). Measurements of the Knight shift, which is linearly proportional to the electron spin polarization, have unambiguously proven the existence of these exotic topological charge excitations. However, this rich probe keeps providing us with more surprising information about the Skyrmions, continuously challenging the theory of QHF. Recent experiments carried out by N. N. Kuzma et al., using optical pumping techniques to enhance the NMR signal of the 2DEG, have focused on the NMR line shape as a function of temperature and filling factor near $`\nu =1/3`$ (which is also a QHF state). The observed free induction decay signals show a large dependence of the electron polarization on filling factor, indicating that charged spin excitations are more important near $`\nu =1/3`$ than previously expected from theory. Near this filling factor, as the temperature is lowered, the Knight shift increases with decreasing temperature (although having a local minimum) until it reaches a saturation level at very low temperatures. The spectral line shape ranges from the motionally narrowed regime, where the polarization shows a sharp peak due to the average polarization seen by each nucleus, to the frozen regime, where the peak is much broader due to the presence of frozen spin textures in the ground state and the nuclei seeing different electron polarizations depending on their location. The interesting regime is the intermediate one where the dynamical time scale of the Skyrmions is comparable to the inverse frequency of the Knight shift. This is the regime which contains the greatest wealth of information about Skyrmion dynamics. The situation is complicated by the fact that the Knight shift varies strongly with position in the $`z`$ direction across the quantum well. Nuclei at the edges of the well see a lower electron density and hence a smaller Knight shift. As a result, these nuclei can still be in the motionally narrowed regime when nuclei in the center of the well are already close to the frozen regime. This subtlety is taken into account in our analysis. Because the spin stiffness at $`\nu =1/3`$ is so small the Skyrmions will be very small and the continuum field theoretic approach will be poorly controlled. We therefore focus on the case of filling factors near $`\nu =1`$ where measurements are currently underway. <sup>,</sup> We present a simple model for the Skyrmion dynamics near $`\nu =1`$ in this paper. We find that the spectral line shape, besides being dependent on the transport relaxation time of the Skyrmions, is very sensitive to the nuclear spin polarization density along the $`z`$ direction and, to a lesser extent, the electron density profile in the z-direction. We organize this paper as follows. In Sec. II we introduce the theoretical background needed for our model calculations. In Sec. III we present our model and results. In Sec. IV we discuss the implications of the results and possible new outlooks on this problem. ## II THEORY In the free induction decay NMR experiments, after the nuclei in the wells are polarized by optical pumping,<sup>,</sup> the nuclear spins are tipped by a $`\pi /2`$ pulse and allowed to precess freely. These spins will precess at the Larmor frequency produced by the local magnetic field. This local magnetic field is composed of the external one plus a contribution from the electron polarization due to the Fermi contact hyperfine coupling which enters the nuclear spin Hamiltonian in the same way as the external magnetic field $$_\mathrm{N}=g_N\mu _N\underset{j}{}𝐒_j\left(𝐇_\mathrm{o}+𝐁_e(𝐑_j)\right),$$ (1) where $`\mu _N`$ is the nuclear magneton, $`𝐒_j`$ is the nuclear spin in units of $`\mathrm{}`$ at position $`𝐑_j`$, $`𝐇_\mathrm{o}`$ is the applied magnetic field, and $`𝐁_e(𝐑_j)(16\pi \mu _B/3)_i𝐒_i^e\delta (𝐫_i𝐑_j)`$ is the effective local magnetic field contribution due to the electronic polarization with $`𝐒_i^e`$ and $`𝐫_i`$ being the spin and position of the $`i`$th electron. The last term in eq. (1) is the one responsible for the observed Knight shift. For the purposes of computing the Knight shift, it is adequate to replace $`𝐁_e`$ by its expectation value $$B_z(𝐑,\nu ,T)|u(𝐑)|^2𝒫(𝐑,\nu ,T),$$ (2) where $`|u(𝐑)|^2`$ is the electron envelope function obtained from a self-consistent local spin-density approximation calculation, and $`𝒫(𝐑,\nu ,T)`$ is the average electron spin polarization at position $`𝐑`$ for a given filling factor and temperature. If we assume that the electron envelope function is only a function of $`z`$ (the growth direction) we can further parameterize the local Knight shift as $`K_s(z)\rho _e(z)\stackrel{~}{K}_s𝒫(𝐑,\nu ,T)`$, where $`\rho _e(z)=𝑑x𝑑y|u(𝐫)|^2`$ is the electron density along the z-direction (normalized to unity at its maximum) and $`\stackrel{~}{K}_s`$ is a constant that can be fitted to the experimental spectra at the lowest temperatures. Next, we must connect these expressions to the observed intensity spectrum $`I(\omega )`$, which is the time Fourier transform of the induced voltage produced in the tipping coil due to the precessing nuclear spins in the quantum wells. In the absence of in-plane spin nuclear decay, for a given nuclear spin, the time evolution of the spin’s expectation value (relative to the evolution with zero Knight shift) is given by $`S_j^+(t)`$ $``$ $`S_{jx}+iS_{jy}`$ (3) $`=`$ $`S_j^+(0)\mathrm{exp}[i{\displaystyle _0^t}𝑑\tau \stackrel{~}{K}_s\rho _e(z_j)𝒫(𝐑_j,\nu ,T,\tau )],`$ (4) where now we have allowed the electronic polarization to vary with time. The induced voltage is proportional to the time derivative of $`\mathrm{Re}\left[S_j^+(t)\right]`$, which is approximately $`\mathrm{Im}\left[i\omega _0S_j^+(t)\right]`$ whenever $`\omega _0\stackrel{~}{K}_s`$, with $`\omega _0`$ being the bare nuclear precession rate. However, for nuclear spins in a solid the in-plane magnetization decays as a gaussian due to nuclear dipole-dipole interactions. This yields the final expression for the intensity $$I(\omega )𝑑𝐫\rho _\mathrm{N}(𝐫)\mathrm{Re}\left[_0^{\mathrm{}}𝑑te^{\sigma ^2t^2/2+i\omega t}S^+(𝐫,t)\right],$$ (5) where $`\rho _\mathrm{N}(𝐫)=_j\delta (𝐫𝐑_𝐣)S_{zj}`$ is the polarization density of nuclear spins (not the number density of nuclei), and $`\sigma =\mathrm{\Delta }\omega /(2\sqrt{2\mathrm{ln}2})`$ with $`\mathrm{\Delta }\omega `$ being the full width half maximum (FWHM) of the unshifted NMR signal. In the samples used in the experiments $`\mathrm{\Delta }\omega =2\pi 3.5\mathrm{kHz}`$, and therefore $`\sigma =9.34\mathrm{msec}^1`$. Any model describing the observed spectra has to contain realistic estimates of $`𝒫(𝐑_j,\nu ,T,t)`$, $`\rho _\mathrm{e}(𝐫)`$, and $`\rho _\mathrm{N}(𝐫)`$. Of these three the first two are the ones that have been studied most extensively. <sup>-</sup> In a previous model presented by Kuzma et al. to describe the transition near $`\nu =1/3`$, a two polarization model for $`𝒫(\nu ,T,t)`$ was used, and $`\rho _e`$ and $`\rho _N`$ were approximated by sinusoidal shapes. Here we attempt to improve upon this model by carefully examining the approximations used for each quantity. At zero temperature $`𝒫(𝐑_j,\nu )`$, the electron polarization, is exactly unity at $`\nu =1`$ as mentioned in section I. However, as the filling factor goes away from from $`\nu =1`$, Skyrmions begin to appear in the ground state of the system. These charged spin texture excitations reduce the polarization locally and a realistic approximation for the shape of the Skyrmion is needed to obtain a reasonable $`𝒫(𝐑_j,\nu )`$. There have been many studies on the shape of the Skyrmions. <sup>-</sup> Most analytical approaches describing the Skyrmion excitations have taken the route of effective field theories such as the modified O(3) non-linear sigma model (NL$`\sigma `$), where a Zeeman term and a Coulomb interaction term are inserted in the classical NL$`\sigma `$ model. The magnetization profile of a Skyrmion obtained from this theory, in the limit of zero Zeeman energy, is given by $`m_z(𝐫)={\displaystyle \frac{r_{}^2\lambda ^2}{r_{}^2+\lambda ^2}},`$ where $`\lambda `$ determines the size of the Skyrmion, and $`r_{}`$ is the projection of $`𝐫`$ on the plane (denoted by $`r`$ henceforth). At larger distances away from the Skyrmion’s center the Zeeman term dominates and the magnetization goes as $`1m_z(𝐫)e^{2\kappa r}/r`$, with $`\kappa =4.4l_B`$ near $`\nu =1`$. This analytical approach has been very successful in explaining qualitatively the physics of the QHF, however it has not been able to predict quantitatively the shape of the Skyrmion excitations at experimentally accessible parameters. In GaAs/Al<sub>x</sub>Ga<sub>1-x</sub>As heterostructures, at around 10 T, the number of spin flips, $`K`$, per unit charge introduced near $`\nu =1`$ is $`34`$. Single Skyrmion microscopic calculations using techniques such as Hartree-Fock (HF), exact diagonalization, and variational wave functions, have been more successful at obtaining quantitative agreement with experiments. These, however, are not as transparent as the classical model in describing the physics behind the excitations. Also, Brey et al. have performed HF calculations of the skyrme crystals formed at filling factors near $`\nu =1`$ which most accurately describe the spin polarization observed in experiments. The failure of the NL$`\sigma `$ model to predict the correct Skyrmion shape for small $`K`$ originates in the truncation of the gradient expansion. The HF calculations are in essence self-consistent mean field calculations with the order of the gradient expansion taken to infinity and hence are more successful at predicting the small Skyrmions that change shape on a much shorter length scale. As described in the next section, we use a phenomenological form for $`m_z(r)`$ fitted to the Hartree Fock calculations. The electron density along the z-direction $`\rho _\mathrm{e}(z)`$, is obtained from the electronic spin-split energy levels of the GaAs heterostructures in the presence of a strong tilted field (as is used in the experiments). These charge distributions are calculated using a local spin-density approximation. The density profiles of the two spin states and two lowest levels, measured with respect to the valence band for a $`30\mathrm{nm}`$-wide GaAs single quantum well, are shown in Fig. 1. The parameters have been chosen to match the experiments done near $`\nu =1`$. The nuclear polarization profile estimate $`\rho _N`$, comes from considering the experimental set up. After a train of rf pulses which destroys the thermally induced nuclear polarization, the sample is radiated with circularly polarized $`\sigma ^+`$ light tuned to the band gap in the well. This excites 100% polarized electrons to the lowest unoccupied level (spin down) and in the recombination process the nuclear system absorbs part of the angular momentum transferred to the sample. Hence, the optical pumping can produce an enhancement in the nuclear polarization by as much as a factor of 100. This optical pumping is necessary, at least at higher temperatures, to make the NMR signal visible. The problem is that, even though the initial excitation process is understood, the recombination process is much more complicated. In such a process, Skyrmions and anti-Skyrmions are being created constantly and this may have an effect on the nuclear polarization profile which is not well understood at present. One may ask why at such low temperatures (1.6 K) and being so deep in the insulating phase of the quantum Hall state one can even observe motional narrowing of the NMR line shapes. We can answer this question in reverse. Given the information from the experiments, what longitudinal resistivity can we infer? The onset of the frozen regime indicates that $`DK_s^1n_{skyr}^1`$ and $`dn/d\mu n/\mathrm{\Delta }E`$, with D being the diffusion constant, $`n_{skyr}`$ and $`n`$ the density of Skyrmions and electrons respectively, and $`\mathrm{\Delta }E`$ the disorder broadening of the Landau level which can be estimated to be at least of the order of $`10`$K and is possibly much larger. This information can be inserted in the Einstein relation for the conductivity $`\sigma _{xx}`$ $`=`$ $`e^2D{\displaystyle \frac{dn}{d\mu }}e^2n_{skyr}^1{\displaystyle \frac{K_sn}{\mathrm{\Delta }E}}`$ (6) $``$ $`{\displaystyle \frac{3\times 10^{12}\mathrm{\Omega }^1}{|\nu 1|}},`$ (7) which gives a lower bound on the inverse conductivity $`1/\sigma _{xx}300\mathrm{G}\mathrm{\Omega }|\nu 1|`$. Hence, we see that the dynamics of the Skyrmions can appear to be fast on the NMR time scale even deep in the insulating regime. ## III THE MODEL AND RESULTS When calculating the spectral intensity from eq. (3) it is useful to first think about the different time scales in the problem and how relevant each one is in calculating the NMR spectral line shape. At the experimental fields used (7.05 T) the bare nuclear precession rate is of order 100 MHz, the extra precession rate created by the electron spins (the Knight shift) is of order 20 KHz, the nuclear spin-lattice relaxation rate ranges from 4 mHz to 45 mHz, and the in-plane spin relaxation rate due to the nuclear dipolar coupling is approximately 3.5 KHz. Hence, in the calculations that follow, we shall ignore the spin-lattice relaxation rate and place our zero of frequency at the bare NMR resonance. Furthermore, since we assume perfect cubic symmetry of the GaAs crystal, we omit any effects on the spectral calculation due to any nuclear quadruple splitting. Rather than doing a full microscopic calculation for $`𝒫(𝐑,\nu ,T)`$ in the presence of disorder, we take a more modest aim and focus on the transition where the spectral line shape goes from the frozen regime to the motionally narrowed regime. For $`|\nu 1|1`$ the ground state of the 2DEG at $`T=0`$ is believed to be in a skyrme square lattice state. In practice, the skyrme lattice is melted in most of the accessible temperature ranges. However, at the low experimental temperatures considered here (T$`<4`$ K), although the long length scale correlations vanish, we expect the short length scale correlations to contain crystal-like features. To model this we introduce a Skyrmion square lattice with the unit cell size given by the appropriate filling factor. We take the magnetization profile of the Skyrmion to be: $`m_z(r)={\displaystyle \frac{r^2\lambda ^2e^{\alpha \frac{r^2}{L^2}}}{r^2+\lambda ^2e^{\alpha \frac{r^2}{L^2}}}},`$ which, by choosing $`\lambda `$ and $`\alpha `$ appropriately, can resemble closely the HF calculations previously done. For $`\nu =0.96`$ we used $`\alpha =5.6`$ and $`\lambda ^2=2.6`$. Although in the HF calculations $`m_z(0)0.6`$ due to the zero point fluctuations, we find that this has little effect on the spectral line shape, and we hence keep our simpler functional form in the model calculations. Also, in any experimental situation, the part of the spectrum due to this small region of fully reversed spins tends to be weak. To model the time dependence of $`𝒫(𝐑,\nu ,T,t)`$, instead of allowing the Skyrmions to undergo correlated thermally induced motion about their lattice points, we make the whole lattice move together to simplify the numerical calculations. Hence, the lattice is only allowed to move collectively in a random walk by performing a jump of average distance $`l`$ with a probability $`dt/\tau _J`$ in the time interval $`t`$ to $`t+dt`$. The diffusion time across a unit cell of size $`L`$, $`\tau _{\mathrm{diff}}=(L^2/l^2)\tau _J`$, is held constant for a given temperature and $`\tau _J`$ and $`l`$ are varied to test the sensitivity of the spectral line shape to the microscopic details of the dynamics. The motionally narrowed regime occurs when $`\tau _{\mathrm{diff}}K_s^1`$ and each nucleus experiences the average electron polarization in the sample. In the limit $`\tau _{\mathrm{diff}}K_s^1`$ the Skyrmions are spatially frozen on the experimental time scale, with random motion being rare. This frozen regime of the spectral line shape, observed at the lowest temperatures, contains information on the actual shape and static distribution of the Skyrmion excitations and the nuclear density profile. The intermediate regime, $`\tau _{\mathrm{diff}}K_s^1`$, occurs when the FWHM of the spectrum reaches its peak value. In this regime we find the greatest sensitivity of the line shape to the choice of $`\tau _\mathrm{J}`$ for a fixed $`\tau _{\mathrm{diff}}`$. The adjustable parameter $`\stackrel{~}{K}_s`$, is fixed by fitting the peak frequency of the spectral line shape to the one observed experimentally in the frozen regime. The temperature is gauged by $`\tau _{\mathrm{diff}}`$ and calibrated by the onset of the motionally narrowed and frozen regimes observed in the experiments. We approximate the nuclear magnetization density by the electronic density from the lowest unoccupied level (E1$``$ in Fig. 1), since the induced polarization due to optical pumping is proportional to the local electronic density of the electrons excited in such process. This gives a much more accurate profile than the ones approximated by simple sinusoidal shapes. One striking result of our model is the high sensitivity of the spectral line shape to the nuclear magnetization density. This sensitivity is illustrated in Fig. 2 where we show different spectral line shapes at $`\nu =1`$ for several nuclear polarization profiles. This result strongly suggest that the details of the line shape will be difficult to understand from first principles without a better microscopic understanding of the optical pumping process and its effect on the nuclear polarization profile, together with other processes that may be affecting this profile. In Fig. 3 we show the spectral line shapes at different diffusion times (temperatures). The maximum FWHM is obtained between 40 and 50 $`\mu `$s. Also, note that $`K_s`$, which corresponds to the maximum in the spectra, decreases monotonically as the temperature (or $`\tau _{\mathrm{diff}}^1`$) increases. It is important to note that at temperatures where the Skyrmion dynamics begin to ‘freeze’ in the NMR time scales, the peak of the spectrum (what is usually understood by the Knight shift) is no longer a good measure of the global average electron polarization, $`\overline{𝒫}d^2r𝒫(r,\nu ,T,t=0)`$. Instead, in order to avoid dynamic effects, one should measure the first moment of the spectra when measuring $`\overline{𝒫}`$, since $$_{\mathrm{}}^{\mathrm{}}𝑑\omega \omega I(\omega )\left[𝑑z\rho _N(z)\rho _e(z)\right]\stackrel{~}{K}_s\overline{𝒫},$$ (8) where here we have $`_{\mathrm{}}^{\mathrm{}}𝑑\omega I(\omega )=1`$ rather than $`I_{\mathrm{max}}=1`$ as used in Fig. 3 and 4. The sensitivity of the line shape to the choice of $`\tau _J`$ at $`\tau _{\mathrm{diff}}4050\mu \mathrm{s}`$ (corresponding to the maximum FWHM) is shown in Fig. 4. This sensitivity, although weak, is completely absent at other $`\tau _{\mathrm{diff}}`$’s in the other regimes. ## IV CONCLUSION Our model illustrates qualitatively the behavior of the NMR spectra as a function of temperature for $`T<4`$ K. It accurately predicts a peak in the FWHM as a function of temperature. It also shows a monotonic increase of the Knight shift with decreasing temperature, reaching a plateau at the lowest temperatures (largest $`\tau _{\mathrm{diff}}`$), however, it does not reproduce the local minimum in the Knight shift as a function of temperature observed near $`\nu =1/3`$. Since the spin stiffness near $`\nu =1/3`$ is very small, we do not expect this model to be valid near such filling factor, where the Skyrmions are small and not well understood as is the case near $`\nu =1`$. We have also demonstrated that a full understanding of the NMR line shape must involve a better understanding of the nonequilibrium nuclear polarization profile. This profile is affected primarily by the optical pumping and possibly by other thermal relaxation processes. We also have shown that to measure the behaviour of the average electron polarization $`\overline{𝒫}`$ at these temperatures, one must measure the first moment of the intensity spectrum rather than its peak. We emphasize that our model does not attempt to calculate the highly quantum-mechanical motion of the Skyrmions. This motion may involve a semi-classical percolation in the case of heavy nonlocalized Skyrmions or variable range hopping in the case of highly localized Skyrmions in a random potential. Our phenomenological model does capture the appropriate time scales in the experiments which should be an essential result of a more sophisticated treatment of the problem. The authors would like to thank René Côté and M. Abolfath for providing the HF data for comparison, and Allan MacDonald, S.E. Barrett, S. Melinte, V. Bayot, and S. Das Sarma for helpful discussions. This work was supported by grant NSF DMR 9714055, by grant INT-9602140, by the Ministry of Education of the Czeck Republic under grant ME-104 and by the Grant Agency of the Czeck Republic under grant 202/98/008J.
no-problem/9908/astro-ph9908293.html
ar5iv
text
# The DIRECT Project: Influence of Blending on the Cepheid Distance Scale. I. Cepheids in M31 ## 1 Introduction As the number of extragalactic Cepheids discovered with HST continues to increase and the value of $`H_0`$ is sought from distances based on these variables (e.g. Saha et al. 1999; Mould et al. 2000), it becomes even more important to understand various possible systematic errors which could affect the extragalactic distance scale. Currently, the most important systematic is a bias in the distance to the Large Magellanic Cloud, which provides the zero-point calibration for the Cepheid distance scale. The LMC distance is very likely significantly shorter than usually assumed (e.g. Udalski 1998; Stanek et al. 2000), but it still might be considered uncertain at the $`10`$% percent level (e.g. Jha et al. 1999). Another possible systematic, the metallicity dependence of the Cepheid Period-Luminosity (PL) relation, is also very much an open issue, with the empirical determinations ranging from 0 to $`0.4magdex^1`$ (Freedman & Madore 1990; Sasselov et al. 1997; Kochanek 1997; Kennicutt et al. 1998). In this paper we investigate a much neglected systematic, that of the influence of blended stellar images on the derived Cepheid distances. Although Cepheids are very bright, $`M_V4`$ at a period of $`10days`$, their images when viewed in distant galaxies are likely to be blended with other nearby, relatively bright stars. We define blending as the close projected association of a Cepheid with one or more intrinsically luminous stars, which can not be detected within the observed point-spread function (PSF) by the photometric analysis (e.g., DAOPHOT, DoPHOT). Such blended stars are mostly other young stars which are physically associated $``$ from actual binary and multiple systems to companions which are not gravitationally bound to the Cepheid. Blending is thus a phenomenon different from crowding or confusion noise; the latter occurs in stellar fields with a crowded and complex background due to the random superposition of stars of different luminosity. In this paper we are concerned with blending due to wide unbound systems. Binary Cepheid companions are well studied (Evans 1992), and do not contribute enough flux to affect Cepheid distances (Madore 1977), due to the obvious constraints of coeval stellar evolution. On the other hand, the association of Cepheids with other luminous stars in wide unbound systems is an unsolved problem in general. While such association, i.e. a strong star-star correlation function, is expected and common to young stars (Harris & Zaritsky 1999), the specific case for Cepheids is unknown. Studies in our Galaxy are very difficult due to the small sample and existing results, though tantalizing, are inconclusive (Evans & Udalski 1994). This could explain the relative neglect of this issue in recent years, but blending had been of concern for early studies of Magellanic Cloud Cepheids (DeYoreo & Karp 1979; Pel, van Genderen & Lub 1981), because even faint B-star blends affect the optical colors of a Cepheid significantly. We investigate the effects of stellar blending on the Cepheid distance scale by studying two Local Group spiral galaxies, M31 and M33. In this paper we concentrate on M31 (Andromeda Galaxy), located at approximately $`R_{M31}=780kpc`$ (e.g. Holland 1998; Stanek & Garnavich 1998) from us. As part of the DIRECT project (e.g. Kaluzny et al. 1998; Stanek et al. 1998) we have collected for this galaxy an extensive data set, finding among other variables 206 Cepheids. We identify some of these Cepheids on archival HST-WFPC2 images and compare them to our ground-based data to estimate the impact of blending on our photometry, taking advantage of their superior resolution – the FWHM on the WFPC2 camera corresponds to $`0.4pc`$ at the distance of M31, compared to $`5pc`$ for the ground-based data. The average FWHM on the DIRECT project ground-based images of M31 is about $`1.5\mathrm{}`$, or $`5pc`$, which corresponds to the HST-WFPC2 resolution of $`0.1\mathrm{}`$ for a galaxy at a distance of $`10Mpc`$. Any luminous star (or several of them) in a volume of that cross section through the disk (at the inclination of the galaxy) could be indistinguishable from the Cepheid and would contribute to its measured flux. As Cepheids are relatively young stars, they reside strictly in the midplanes of the disks of spiral galaxies. The archival HST-WFPC2 study of M31 Cepheids from our project DIRECT was undertaken to improve our distance determination to the galaxy. The preliminary indirect (via LMC) Cepheid distance we obtained (Kaluzny et al. 1998; Sasselov et al. 1998) was practically the same as the Cepheid distance by Freedman & Madore (1990) of $`R_{M31}=770\pm 25kpc`$. Our findings of blending now indicate that these distance estimates should be corrected upward. We describe the ground-based and HST data and the applied reduction procedures in Section 2. In Section 3 we discuss the task of identifying Cepheids on HST WFPC2 images. In Section 4 we present the blending catalog of Cepheids and discuss it in Section 5. In Section 6 we describe the effect of blending on the light and color curves of Cepheids and how these curves can be used to detect blends. In Section 7 we discuss the effectiveness of some of the blend analysis methods encountered in the literature. The concluding remarks are to be found in Section 8. ## 2 Observations and Data Reduction ### 2.1 Ground-based Data The ground-based data were obtained as part of the DIRECT project between September 1996 and October 1997 during 95 full/partial nights on the F. L. Whipple Observatory 1.2 m telescope and 36 full nights on the Michigan-Dartmouth-MIT 1.3 m telescope. Six $`11^{}\times 11^{}`$ fields with a scale of 0.32 ″/pixel were monitored: four of them (A–D) concentrated on the rich spiral arm in the northeast part of M31, one (E) close to the bulge of M31 and one (F) containing the giant star formation region known as NGC 206. Fields A–D and F have been reduced and the $`BVI`$ photometry of Cepheid variables published in Stanek et al. 1998 (hereafter, Paper II), Kaluzny et al. 1998 (Paper I), Stanek et al. 1999 (Paper III), Kaluzny et al. 1999 (Paper IV) and Mochejska et al. 1999 (Paper V), respectively. The applied reduction, calibration and variable selection procedures are discussed therein, particularly in Paper I, where full details are provided. A total of 206 Cepheids were found: 43 in field A, 38 in B, 35 in C, 38 in D and 52 in field F. ### 2.2 HST data The archival HST-WFPC2 data used in this paper were retrieved from the Hubble Data Archive. We selected images overlapping our M31 ground-based data, taken in filters F336W (roughly $`U`$), F439W, F450W ($`B`$), F555W, F606W ($`V`$) and F814W ($`I`$). The pixel scales of the Wide Field (WF) and Planetary Camera (PC) chips are 0.0996 and 0.0455 ″/pixel, respectively. The full list of exposures is provided in Table 1, along with the proposal ID, dataset name, equatorial coordinates of the frame centers, filter and exposure time information. The data we obtained had already passed through the standard preliminary processing and calibration procedures prior to its placement in the Archive. The standard pipeline calibration is fully described in the HST Data Handbook. The first two steps in our reduction procedure were to mark the bad pixels on the images and to compensate for the fact that pixels on the edges and corners of the CCD receive fewer photons due to the geometric distortion in the WFPC2 optics. For each image a mask was created from the data quality file retrieved from the Archive and a vignetting mask generated by Stetson (1998) and then used to mark bad pixels and vignetted regions. To restore the integrity of the flux measurements the images were multiplied by a pixel-area map, originally created by Holtzman et al. (1995) and renormalized to the median pixel area on each chip by Stetson (1998). These tasks were accomplished under IRAF<sup>1</sup><sup>1</sup>1IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the NSF.. In the next step of the reduction procedure pairs (or multiplets) of images were selected, taken in the same filters, having identical center coordinates as well as similar exposure times. The crrej task under IRAF was used to combine the images and remove cosmic rays. When multiple images for a given field were not available, single images were used for photometry. The photometry was extracted by means of the DAOPHOT/ALLSTAR package (Stetson 1987, 1992). Stars were identified using the FIND subroutine and aperture photometry was done on them with the PHOT subroutine. We used the point-spread functions (PSF) derived individually for each filter and chip of the WFPC2 camera, kindly provided to us by P. B. Stetson (private communication) to obtain the ALLSTAR profile photometry. After the initial ALLSTAR run, FIND was ran again on the star subtracted images, to identify stars that were missed on the first pass. Aperture photometry was obtained for them with PHOT, followed by profile photometry with ALLSTAR. The two star lists were merged and used as input to ALLSTAR to obtain the final photometry. It should be noted that the HST photometry has not been calibrated to any standard system and therefore instrumental magnitudes are used throughout this paper. This has, however, no bearing on the results presented in this paper, since they are strictly based on differential photometry. ## 3 The Identification of DIRECT Cepheids in HST Data The preliminary identification of DIRECT Cepheids on the HST frames was accomplished by a visual comparison of the HST data matched via World Coordinate System (WCS) information to our ground-based $`V`$ template image for each field, using SAOimage ds9<sup>2</sup><sup>2</sup>2SAOimage ds9 was developed under a grant from NASA’s Applied Information System Research Program (NAG5-3996), with support from the Chandra Science Center (NAS8-39073). The WCS information for the ground-based template image header was obtained with a program written by Mink (1997; 1999) by matching stars from the image to the USNO A2.0 Catalog stars (Monet et al. 1996). We have matched a total of 22 Cepheids to HST data: six in our field B, four in C (two of them on two different overlapping fields) and 12 in field F (one found on two different fields). This constitutes $`18\%`$ of the 125 Cepheids in fields B, C and F. In some cases, when a Cepheid which appeared to be a single star on the $`V`$ template was resolved into multiple stars on an HST image, it was difficult to distinguish which star was the Cepheid on the basis of the template image alone. Two such cases are shown in Figure 1. The images plotted in the left panels are taken from the $`V`$ template, created by averaging together two images with the best seeing (FWHM $`1\mathrm{}`$). The image in the upper right panel was taken with the PC chip of the WFPC2 camera while the one in the lower right panel comes from a WF chip. The inset has different IRAF z-scale limits to show the two blended stars separately. The circles, centered on the Cepheids, have a radius of $`0.75\mathrm{}`$. The Cepheid V4125 D31F is shown in the upper panels of the Figure 1. On the ground-based template it appears as a single star, while on the PC chip it is resolved into two objects, one 3 magnitudes dimmer than the other. A third star appears below and to the right, just outside the circle, fainter by almost 2 magnitudes from the brighter component of the doublet. That star was not identified on the ground-based template image, but after subtracting the Cepheid a brighter spot is visible at that location. The lower panels show the case of V2203 D31F, where the Cepheid, appearing as a single star on the ground $`V`$ template image, is resolved by HST into two stars, differing by one magnitude in brightness. To help confirm the Cepheid nature of the selected objects, instrumental color-magnitude diagrams (CMDs) were constructed from HST data, whenever photometry in two bands was available. A few representative CMDs are shown in Figure 2, ($`v_{F555W},v_{F555W}i_{F814W}`$) in the left panels and ($`v_{F555W},b_{F439W}v_{F555W}`$) in the right panels. The Cepheids are denoted by circles and their companions by squares. Stars from the same chip are plotted in the background for reference. The upper left panel shows a case where there is a substantial contribution of flux from a red giant companion ($`(\frac{f}{f_C})_V=34\%`$, $`(\frac{f}{f_C})_I=67\%`$) which is seen at a distance of $`0.2\mathrm{}`$ from the Cepheid V1893 D31F. The other companion is a blue star at a distance of $`0.4\mathrm{}`$ which appears to be located at the red edge of the main sequence (also in the $`v_{F555W},u_{F336W}v_{F555W}`$ CMD which is not shown). The lower left panel presents a case where the Cepheid V2203 D31F has a luminous blue main sequence companion with $`(\frac{f}{f_C})_V=40\%`$ and $`(\frac{f}{f_C})_I=17\%`$ which are separated by $`0.3\mathrm{}`$. The upper right panel shows the Cepheid V7184 D31B with two blue main sequence companions, the more luminous one, located $`0.4\mathrm{}`$ away, contributing $`45\%`$ as much light as the Cepheid in the $`B`$ band and $`23\%`$ in $`V`$. A typical situation where none of the companions (in this case one) had $`b`$-band photometry (see Table 2) is illustrated in the lower right panel of Figure 2, where the Cepheid V4954 D31B is shown. ## 4 M31 Blending Catalog ### 4.1 The Catalog The criterion for blending was determined empirically, by careful examination of ground-based and HST data for the same Cepheids. We consider a star to be blended with the Cepheid if it is located at a distance of less than $`0.75\mathrm{}`$ from it and is not detected by DAOPHOT in our ground-based images. The choice of maximum distance was motivated by the typical full width at half maximum (FWHM) in our ground-based images ($`1.5\mathrm{}`$). Due to the relatively small number of Cepheids having HST data, it was deemed worthwhile to examine them on the HST images in detail to see whether all of the companions had been identified by DAOPHOT and to check for the possibility of false detections (cosmic rays in case of single images, bad columns, etc.). In the latter case the object was removed from the list. Not much could be done in the former case, although that proved not to be a problem and only in a few of the 55 examined images there were very faint companions which eluded detection. It must be noted that the FWHM of the PSF on the WF chip is of the order of 1 pixel and so the probability of the detection of a faint star may depend on the way its light is distributed over the pixels. One exception, however, is V1893 D31F, which DAOPHOT failed to recognize as a star on one of the single F606W filter images, along with its companion 2 pixels ($`0.2\mathrm{}`$) away. No problem was encountered on the other three combined images. We estimate that our data is fairly complete for companions contributing at least $`4\%`$ of the flux of the Cepheid in the V band (filters F555W and F606W). We have used this somewhat arbitrary cutoff in evaluating the sum $`S_F`$ of all flux contributions in filter $`F`$ normalized to the flux of the Cepheid: $$S_F=\underset{i=1}{\overset{N_F}{}}\frac{f_i}{f_C}$$ (1) where $`f_i`$ is the flux of the i-th companion, $`f_C`$ the flux of the Cepheid on the HST image and $`N_F`$ the total number of companions. In Table 2 we present the results for each of the 22 Cepheids found in the HST images: the name, the mean $`V`$, $`I`$ and $`B`$ magnitudes taken from Papers II, III and V, the number of companions $`N_F`$ and their total flux contribution $`S_F`$ in the $`V`$, $`I`$, $`B`$ and $`U`$ bands respectively. Unless noted otherwise, the $`V`$, as in $`N_V`$ and $`S_V`$, refers to filter F555W, $`I`$ to F814W, $`B`$ to F439W and $`U`$ to F336W. For the two Cepheids identified on two HST images, the average values of $`S_F`$ are listed. The catalog of Cepheid blending in M31 is illustrated in the following two figures. Figure 3 shows a histogram of the number of Cepheids $`N_V`$ as a function of $`S_V`$ (Eq. 1), the sum of flux contributions from its companions in the $`V`$ band (filters F555W and F606W), normalized to the flux of the Cepheid. The upper left and right panels and the lower left panel show the histograms for fields B, C and F, respectively. In the lower right panel the combined data is shown. The width of each bin is 0.1 and the first one starts at 0. For further discussion on Figure 3 see §5. A diagram showing the flux contribution from companions $`S_V`$ as a function of the $`V`$ magnitude of the Cepheid obtained from ground-based data (Papers II, III and V) is presented in Figure 4. Field B Cepheids are denoted by squares, field C by triangles and field F by circles. ### 4.2 The Blended Cepheids and their Environments The first step in using the derived blending is to see what is the effect on the Cepheid period-luminosity (P-L) relation. Figure 5 illustrates the influence of blending on the location of the Cepheids on the P-L diagram. The solid circles show the original locations of the Cepheids, based on their mean $`V`$ magnitudes obtained from the ground-based data. The arrows illustrate the shift in $`V`$ when the effects of blending are taken into account. It has to be kept in mind that neither the differential reddening in M31 nor the random-phase nature of the HST data have been accounted for, thus there is significant scatter in the diagram. However, the direction and the magnitude of the effect is obvious. A pattern in the spatial distribution of Cepheid blending in M31 could show us a correlation between blending and crowding. We find no such correlation – Cepheids in environments of different surface brightness show roughly the same frequency and amount of blending, thus distinguishing the phenomenon of blending from crowding. In Figure 6 the blending parameter $`S_V`$ is plotted as a function of the surface brightness around the Cepheid for which it was determined. The surface brightness was taken to be the mode within a 20 pixel radius on the DIRECT template frames. We used a bin width of 1 ADU for the histogram to compute the value of the mode, smoothed with a flat-topped rectangular kernel (boxcar) filter, 11 units in length. After correcting for the sky level, the instrumental surface brightness values were converted to mag/□″using 12-17 fairly bright isolated reference stars with known standard magnitudes by means of the following formula: $$SB[\mathrm{mag}/\mathrm{}\mathrm{}]=m_{ref}+2.5\mathrm{log}(s^2I_{ref}/I_{SB})$$ where $`I_{ref}`$ and $`m_{ref}`$ are the flux in ADU/□ pixel and the corresponding magnitude of the reference star, $`I_{SB}`$ is the surface brightness expressed in ADU/□ pixel and $`s`$ is the pixel scale of the chip. The $`rms`$ scatter around the value of the average ranged from 0.025 to 0.061 mag/□″. In order to give these numbers some perspective, we have computed surface brightness around Cepheids in two galaxies which straddle their range, NGC 2541 and NGC 4535 (Fig. 7). We find that the Cepheids we discover in M31 reside in environments of surface brightness typical of spiral galaxies, despite the high inclination of M31. The HST data for these galaxies, observed as part of the HST Key Project on the Extragalactic Distance Scale (Ferrarese et al. 1998, Macri et al. 1999), were obtained from the Hubble Data Archive. Since we were unable to estimate the sky level for those images, we tried to choose epochs where the sky level would be the lowest. We chose the 1995 Nov 20 epoch for NGC 3541 and the 1996 May 24 epoch for NGC 4535. The surface brightness values were computed using the same method as for our M31 Cepheids and were re-expressed in units of mag/□″using the zeropoints provided in the HST Data Handbook. ## 5 Discussion of Blending Properties Before attempting to draw any far-reaching conclusions from Figs. 3-7 or the catalog itself (Tab. 2), we must stress that they are based on a fairly small sample of Cepheids and therefore subject to statistical uncertainties. An illustration of this effect is the small number of counts in the second bin of the combined histogram (Fig. 3), compared to the third and, especially, the first bin, which could be an artifact of small number statistics. An encouraging fact is that the distributions in separate fields, especially B and F, appear to be similar to a substantial degree in their overall shape and spanned range of $`S_V`$. In the combined histogram there is a decreasing trend in the number of cases $`N_V`$ with increasing flux contribution from blending, $`S_V`$. The diagram in Fig. 4 conveys some of the information contained in the previous figure – a lack of cases with $`0.12S_V0.23`$ is readily apparent, which, as mentioned above, is believed to result from the fairly small size of the sample. If we attribute the gap to statistical effects, then Cepheids fainter than $`V20`$ appear to populate a wide range of $`S_V`$, from 0 up to approximately 0.35, with one object at $`S_V0.47`$. Furthermore, it is observed that in none of the fields do Cepheids show a tendency to favor any particular value of $`S_V`$, as also apparent from Figure 3. Another striking feature of the diagram is that all of the three Cepheids brighter than $`V20`$ have $`S_V0.4`$, although it has to be emphasized that this sample is much too small to draw any definite conclusions. One possible explanation may be ventured, however, which may hold true in these particular cases, but does not have to be the rule for the whole population of bright Cepheids. The most luminous Cepheids are the most massive ones and, hence, the youngest. The group or association of stars in which they had formed will have had less time to disperse. This may manifest itself in a somewhat higher spatial density of stars and thus would increase the probability of blending. Figure 6 shows the blending parameter $`S_V`$ of the M31 Cepheids as a function of the local surface brightness. As there is a gap at $`S_V0.2`$, the separation into blended ($`S_V>0.25`$) and unblended Cepheids ($`S_V<0.15`$) is readily discernible in the diagram. No correlation between the blending parameter $`S_V`$ and the underlying surface brightness is apparent. To put the surface brightness values around the M31 Cepheids into perspective, we plot the surface brightness values computed for the NGC 4535 and NGC 2541 Cepheids (Fig. 7). The range of surface brightness of NGC 2541 falls below the range of our M31 Cepheids. There is a 26% overlap between the NGC 4535 and M31 data – 13 out of the 50 Cepheids in NGC 4535 are located in regions with surface brightness in the range covered by our M31 Cepheids. It should be kept in mind that our sample may also be affected to some extent by selection effects: for example, 11 of our Cepheids are located in the vicinity of NGC 206, the giant star-forming region in M31. Additionally, M31 is observed at a rather high inclination angle, which is not a typical situation for most galaxies searched for Cepheids. An opportunity to study the effects of blending with a larger sample of $`100`$ Cepheids, in a more face-on system, will present itself in the next paper on Cepheids in M33 (Mochejska et al. 2000). We will also discuss there the influence of blending on the observed colors of Cepheids. ## 6 Blending and the Light Curves The luminosity variations through the pulsation cycle of a Cepheid seen in the optical are due primarily to changes in temperature. The monochromatic surface brightness changes with temperature and is additionally quite dependent on wavelength. Of course, the Cepheid also expands and contracts $``$ its area varies. This contributes to the total light variation in a wavelength independent way. The radius variation in a typical Cepheid is out of phase by about 90 with respect to the surface brightness variation. Therefore a comparison between color (proxy for temperature variation) and magnitude (proxy for luminosity variation, due to both radius and temperature change) would produce a loop in the CMD plane. These loops, i.e. the tight correlation between color and luminosity, have had a limited use in the past as part of the Cepheid PLC calibration (Fernie 1964), or to transform away the temperature-induced variation (Madore 1985; Moffett & Barnes 1986). The characterization of Cepheid companions had been traditionally done by using the similar loops in the color-color plane of optical bands (Stobie 1970). The color-color loop arises from the phase shift introduced between the pair of color-magnitude relations when a companion of different temperature (i.e. color) is added. Blue or red companions distort the color-color loops in very different and quantifiable ways (Madore 1977; DeYoreo & Karp 1979) and can be very effective in deriving the amount of and color of contamination in a Cepheids flux. In order to understand the wavelength-dependent effect of a blue or red companion on the light curves (and loops) of a Cepheid, one should bear in mind that the contaminant flux is most prominent during the minimum in the Cepheid lightcurve, and gradually becomes insignificant as the Cepheid brightens up during its cycle. The addition of a red companion to the flux of a Cepheid has the following effects on its optical light curves: (1) the light curve exhibits a flatter minimum (due to the added flux); (2) the color curve has a deeper minimum (due to the added red flux); and (3) the asymmetry in the color curve decreases. The addition of a blue companion to the flux of a Cepheid has the following effects on its optical light curves: (1) the light curve exhibits a flatter minimum (as above); (2) the color curve has a flatter minimum (due to the added blue flux); and (3) the asymmetry in the color curve increases. A typical example is the Cepheid V7184 D31B (Paper I), which has blue blends (Table 2) and whose CM loop is shown in Fig. 8. The effect on its loop is predictable given the above discussion: as the Cepheid fades and becomes naturally redder, the added constant blue flux of the blends becomes more prominent and diverts the lower part of the loop to the blue (left in Fig. 8). This diversion could result eventually in a complete reversion of the naturally clockwise trajectory of the loop. If the data is incomplete (there are two gaps in the light curves of V7184), the blue blending is still detectable by the steeper slope of the loop. These changes on the color-magnitude plane, together with a fit of the mean colors of the Cepheid to extinction laws, can be a powerful means to find and characterize blended Cepheids in a sample like our DIRECT project one. The necessary requirement is $``$ good, well-sampled light curves in at least two bands. Therefore we were able to find blended Cepheids like V7184 in M31B and exclude them from our analysis without the benefit of HST images, but just using our light curves (Paper I). However, the technique is limited to the cases with strongest blending – a nonlinear fitting routine by Rebel (1998) applied to the M31B data still did not detect 2/3 of the blended Cepheids we found in this paper. In other words, the naive technique of visual inspection for ”flat-bottom” light curves is completely inadequate for detecting 10%-20% Cepheid blending, which is dominating our findings in M31. ## 7 Blend Analysis by the HST Key Project in the Galaxy NGC 2541 The influence of blending has been largely neglected in many recent galaxy distance determinations based on Cepheids. Few attempts have been made to deal with this problem. On one of the galaxies observed by the HST Key Project on the Extragalactic Distance Scale, NGC 2541 (Ferrarese et al. 1998), several criteria were used to reject Cepheids which may suffer from crowding. We have applied them to our sample of 22 Cepheids to estimate their effectiveness in detecting blending. As we discussed already in §1, blending and crowding are different phenomena, and judging from §5 they do not seem much correlated in M31 either, but it would be helpful if these criteria could detect blending. One of the proposed tests (Ferrarese et al. 1998) involves identifying Cepheids that have companions not resolved by profile photometry programs by looking at their photometric errors. The underlying idea is that the photometric error would be higher than in case of an isolated star of the same magnitude due to a poorer PSF fit resulting from the unresolved companion. We have constructed diagrams showing the ALLSTAR photometric error $`\sigma _V`$ as a function of the ALLSTAR $`V`$ magnitude, similar to the one presented in Fig. 8 of Ferrarese et al. (1998). The diagram for field F is shown in Figure 9. The Cepheids are plotted as solid dots surrounded by circles with their size proportional to the amount of blending $`S_V`$ (no outside circle means $`S_V=0`$). The $`V`$ magnitudes and the errors are taken from the field F template image (FWHM$`1\mathrm{}`$), as are the stars plotted in the background. Upon examining Figure 9 it is apparent that the photometric errors $`\sigma _V`$ do not show a correlation with the amount of light coming from companions $`S_V`$. One such case is V2320 D31F which is an unblended star with an error almost twice as large than most stars of similar magnitude. Another example is V3289 D31F with $`S_V=0.11`$, which has a photometric error larger than two other, more blended Cepheids of similar $`V`$ magnitude: V1549 D31F with $`S_V=0.28`$ and V3550 D31F with $`S_V=0.47`$. Most of the Cepheids lie within the most densely populated areas of the diagram, with V2320 D31F and V2203 D31F being the most clear outliers. A similar photometric error distribution, uncorrelated with $`S_V`$, is also seen in fields B and C. In conclusion, the presence of close companions did not manifest itself in the form of a poorer PSF fit and, hence, a larger photometric error in the studied sample. Another fact which should be kept in mind is that our ground-based PSF is much better sampled than the HST PSF (comparing to the respective FWHMs), thus it should be more sensitive to any deformations resulting from the superposition of two or more stars. Another proposed method of rejecting blended Cepheids is to remove all stars having companions contributing more than $`50\%`$ of the total light within a radius of two pixels. Taking into account the fact that the two pixels on the HST WF chips correspond to twice the FWHM, we have conducted a similar test on our ground-based $`V`$ template data for the Cepheids. Only one of our Cepheids, V5646 D31B with $`S_V=0.05`$ would be rejected from our sample on the basis of this criterion. This test also failed to identify the blended Cepheids in our sample. A third proposed test for blending is to check if the Cepheid lies on the instability strip in the CMD. The usual color range spanned by the final sample of Cepheids discovered with HST is about one magnitude, for example $`0.4VI1.4`$ in NGC 2541 (Ferrarese et al. 1998), $`0.4VI1.5`$ in NGC 4639 (Saha et al. 1997), $`0.6VI1.5`$ in NGC 4535 (Macri et al. 1999). The examination of a ground-based CMD for the 12 field F Cepheids yields a $`VI`$ color span of also about one magnitude ($`0.6VI1.6`$). No clear correlation of the color with the $`S_V`$ of the Cepheid is seen in the diagram, as unblended Cepheids also exhibit a large color scatter, from $`VI=0.73`$ for V2320 D31F to 1.59 for V3860 D31F, most likely due to differential reddening. In the case of our sample of 22 Cepheids, the tests were not successful in rejecting the blends. These tests may be helpful in some cases of blending, when the two stars are almost resolved or when the companion has a larger flux contribution than the maximum value of $`S_V=0.47`$ present in our sample. But a contamination of $`S_V=0.47`$ has a large impact on the photometry of the Cepheid, which actually would be dimmer by almost one third from the measured brightness. ## 8 Conclusions For our sample of 22 Cepheids with both ground-based and HST data, we find that the mean $`V`$-band flux contribution from companions unresolved in the ground-based images, $`S_V`$, is about $`19\%`$ of the flux of the Cepheid, while the median $`S_V`$ is 12%. This shows that blending could potentially be a substantial source of error in the Cepheid distance scale, as the distance derived from our ground-based photometry for this admittedly small Cepheid sample would be systematically underestimated by $`9\%`$ (for the mean $`S_V`$) or $`6\%`$ (for the median $`S_V`$). This is to be compared to the current Cepheid distance to M31 (which is subject to the LMC distance uncertainty) of $`R_{M31}=770\pm 25kpc`$ (Freedman & Madore 1990; Kaluzny et al. 1998; Sasselov et al. 1998). The Cepheid photometry by Freedman & Madore (1990) was from the Canada-France-Hawaii telescope (average seeing 1.0″, Freedman, Wilson, & Madore 1991). Our findings of blending here require that these distance estimates to M31 be corrected upward by about 9%. Blending becomes even more severe when we consider galaxies at a distance of $`25Mpc`$, i.e. at the edge of what can be currently observed with HST. The $`0.1\mathrm{}`$ FWHM on the WF chips of the WFPC2 camera would span a linear distance of $`12pc`$ in such a galaxy, which corresponds to $`3.2\mathrm{}`$ at the distance of M31. In order to obtain an estimate as to the degree of contamination caused by blending at a distance of $`25Mpc`$, we have summed the contributions of all Cepheid companions within a diameter of $`3.2\mathrm{}`$ in our HST data for M31. The result is presented in Figure 10 in the form of a cumulative probability distribution of $`S_V`$. The diagram shows that $`50\%`$ of our Cepheids have $`S_V>0.4`$, though some of the Cepheids with a very high degree of contamination would probably elude detection. This indicates that blending will very likely introduce a significant contamination to Cepheid photometry at such distances and resolution. As the result of blending with other unresolved stars, the Cepheids appear brighter than they really are when observed in distant galaxies with HST. As we compare them with mostly unblended LMC Cepheids, this leads to systematically low distances to galaxies observed with the HST, and therefore to systematically high estimates of $`H_0`$. The sign of the blending effect on the $`H_0`$ is opposite to that caused by the lower LMC distance (e.g. Udalski 1998; Stanek et al. 2000) and might be of comparable value, as discussed in this paper. It should be stressed that blending is a factor which contributes in only one direction, and therefore it will not average out when a large sample of galaxies is considered. One obvious solution to the problem of blending would be to obtain data with better resolution. Such an opportunity will be available after the launch of the Next Generation Space Telescope, scheduled for 2008. An alternative approach would be to determine the amount and color of flux contamination for the Cepheids by the analysis of their light curves and/or color-color loops. The requirement, however, is to have good quality, well-sampled light curves in at least two bands, which is not the case for HST Cepheids. However, even with our good light curves it is difficult to detect Cepheids with blending of $`S_V<0.20`$. It must be stressed that the images taken with the HST WFPC2 camera, despite having a resolution $`>10`$ times better than our ground-based images (FWHM $`0.4pc`$ vs. $`5pc`$), will still leave some blends unresolved, including physical companions (e.g. Evans 1992). Thus, the $`V`$-band flux contribution from Cepheid companions derived in this paper sets only the lower limit on the true influence of blending on the Cepheid photometry. Bohdan Paczyński, Andrzej Udalski and Thierry Forveille have provided us with helpful comments on the manuscript. We would like to thank Janusz Kaluzny for providing us with his database management codes, Peter Stetson for the HST WFPC2 point-spread functions and Doug Mink for assistance in automating image astrometry. This work was partially based on observations with the NASA/ESA Hubble Space Telescope, obtained from the data Archive at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract No. NAS5-26555. Support for this work was provided by NASA through Grant AR-08354.02-97A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. BJM was supported by the Polish KBN grants 2P03D00317 to Janusz Kaluzny and 2P03D01416 to Grzegorz Pojmański. DDS acknowledges support from the Alfred P. Sloan Foundation. KZS was supported by the Harvard-Smithsonian Center for Astrophysics Fellowship.
no-problem/9908/cond-mat9908378.html
ar5iv
text
# Supercurrent Stability in a Quasi-1D Weakly Interacting Bose Gas ## Abstract We discuss a possibility of observing superfluid phenomena in a quasi-1D weakly interacting Bose gas at finite temperatures. The weakness of interaction in combination with generic properties of 1D liquids can result in a situation when relaxational time of supercurrent is essentially larger than the time of experimental observation, and the behavior of the system is indistinguishable from that of a genuine superfluid. PACS numbers: 03.75.Fi, 05.30.Jp, 67.65.+z The discovery of Bose-Einstein condensation in ultra-cold dilute gases has opened up a unique possibility for the study of fine aspects of superfluidity, including the specifics of lower dimensionality. Primarily, the new opportunities come from the availability of isolated systems, free of contact with walls or any sort of substrate. A special interest is associated with the prospects of creating one-dimensional systems. Actually, one deals here with a quasi-1D situation, when the configuration of a trap yields toroidal spatial particle distribution with small transverse localization size $`l_{}`$. At low temperature, and small interaction energy per particle, $`ϵ_0`$, as compared to the transverse-motion level spacing, the system kinematics and collective dynamics are purely one-dimensional, though the value of the effective interaction is determined by 3D pair collisions (see below). Strictly speaking, the 1D superfluidity cannot exist at $`T0`$. - At finite temperature the relaxational time for supercurrent is finite, i.e., independent of the system size in the limit $`L\mathrm{}`$. Nevertheless, as will be shown below, analyzing the situation with a dilute gas one finds that the relaxation rate of the supercurrent is exponentially reduced with decreasing interaction, and that in a quasi-1D weakly interacting system this can lead to a new quality when at a finite temperature the lifetime of a supercurrent state considerably exceeds the lifetime of the isolated gas sample. We start with rendering basic notions and results for the low-energy properties of 1D (super)fluids. First, we remind that in 1D there exists a condition on superfluid density, $`n_s`$, and compressibility, $`\kappa `$, determining whether the superfluid groundstate is stable against either single impurity, or disorder, or an external commensurate potential. This condition requires that the dimensionless parameter ($`m`$ is the particle mass, $`\mathrm{}`$=1) $$K=\frac{1}{\pi }\sqrt{\frac{m}{\kappa n_s}}\frac{cm}{\pi n_s}$$ (1) be less than corresponding critical value $`K_c`$ (here $`c=\sqrt{n_s/m\kappa }`$ is the sound velocity). For the most important for our purposes cases of single impurity and random potential the critical values of $`K`$ are $`K_c=1`$ and $`K_c=2/3`$, respectively . As we will see below, the case of quasi-1D weakly interacting Bose gas corresponds to $`K1`$, so that the stability of the superfluid groundstate is guaranteed. The low-energy spectrum of any one-component 1D system in the superfluid phase is exhausted by bosonic sound-like excitations and a collective state supporting non-zero supercurrent, $`J_I=2\pi n_sI/mL`$, labeled by an integer index $`I`$ (we consider a system of ring geometry) $$H_0=\underset{q0}{}\omega _qb_q^{}b_q+\underset{I}{}E_Id_I^{}d_I,$$ (2) where $`b_q^{}`$ creates a boson with energy $`\omega _q=cq`$ and momentum $`q=2\pi k/L`$, where $`k`$ is integer. We have used a standard short-hand notation to specify the current state \- obviously $`d_I^{}d_I`$ is either $`1`$ or $`0`$ and $`_Id_I^{}d_I=1`$. The energy of the current state is given by $$E_I=\frac{2\pi ^2n_s}{mL}I^2.$$ (3) Note, that although the energy of the current state scales like $`1/L`$, the momentum associated with it is system-size independent, $`P_I=2\pi n_sI`$. Within the long-wavelength effective Hamiltonian \[describing the superfluid state in terms of the phase field\], $`H=𝑑x[(n_s/2)(d\mathrm{\Phi }/dx)^2+(\kappa /2)\dot{\mathrm{\Phi }}^2]`$, where the original Bose field is written as $`\mathrm{\Psi }=\sqrt{n}e^{i\mathrm{\Phi }}`$, bosonic excitations and current states (sometimes called “zero mode” states ) correspond to the following decomposition $`\mathrm{\Phi }=\varphi (b)+2\pi Ix/L,{\displaystyle (d\varphi /dx)𝑑x}=0,`$ where $`x`$ is a coordinate along the circle. Thus $`I`$ is nothing but a phase field circulation $$I=\frac{1}{2\pi }(d\mathrm{\Phi }/dx)𝑑x.$$ (4) The very notion of the superfluid phase means that $`I`$ is a well-defined quantum number even when rotational invariance is violated. In higher dimensions at $`T<T_c`$ relaxation times $`\tau _I`$ are extremely long with exponential dependence on the system size, and for all relevant experimental time-scales, $`\tau _{\mathrm{exp}}`$, the condition $`\tau _{\mathrm{exp}}\tau _I`$ is satisfied. Special arrangements are to be made to shorten $`\tau _I`$ (e.g., weak-link or tunnel-junction systems). To find relaxation times in 1D we may use the transition Hamiltonian which gives explicitly transition amplitudes between states $`|I,\{N_q\}`$ and $`|I^{},\{N_q^{}\}`$ for arbitrary rotational invariance breaking terms. For the case of a local perturbation $`V(x)`$ with spatial dimension smaller than the correlation radius $`1/mc`$, which without loss of generality may be replaced by $`V(x)V_0\delta (xx_0),`$ the transition Hamiltonian acquires a form $$H_{\text{int}}=gV_0n\underset{II^{}}{}d_I^{}d_I^{}e^{i2\pi n(II^{})x_0}\mathrm{\Lambda }_{II^{}}(b)+h.c..$$ (5) Polaronic-type exponential operator $`\mathrm{\Lambda }(b)`$ in this expression equals to $$\mathrm{\Lambda }_{II^{}}(b)=\mathrm{exp}\left\{\frac{i(II^{})}{\sqrt{K}}\underset{q0}{}\left(\frac{2\pi }{L|q|}\right)^{1/2}\mathrm{sgn}(q)(b_qb_q^{})e^{iqx_0}\right\}.$$ (6) The exact value of the numeric coefficient $`g`$ in Eq. (5) is outside the scope of the long-wavelength treatment used to derive $`H_{int}`$ (it is of order unity for weakly-interacting Fermi gas). Formally, the problem has reduced to the problem of particle dynamics on a 1D lattice with Ohmic coupling to the oscillator bath environment (see, e.g., review ) after polaronic on-site transformation. The dimensionless Ohmic coupling parameter $$\alpha =(II^{})^2/K$$ (7) is large in our case, $`\alpha >1`$, and “particle”/current dynamics is incoherent at any finite temperature. We may then immediately utilize the well-known expression for the transition probability between states $`I`$ and $`I^{}`$ which derives from the Golden-rule expression for $`H_{\text{int}}`$ (see, e.g., Ref. ) $$W=|gV_0n|^2e^Z\frac{2\sqrt{\pi }\mathrm{\Omega }}{\xi _{II^{}}^2+\mathrm{\Omega }^2}\frac{|\mathrm{\Gamma }[1+\alpha +i\xi _{II^{}}/(2\pi T)]|^2}{\mathrm{\Gamma }[1+\alpha ]\mathrm{\Gamma }[1/2+\alpha ]}e^{\xi _{II^{}}/2T},$$ (8) where $$Z=2\alpha \mathrm{ln}\left(\frac{\gamma ϵ_0}{2\pi T}\right),$$ (9) $$\mathrm{\Omega }=2\pi \alpha T,$$ (10) $$\xi _{II^{}}=E_IE_I^{},$$ (11) $`\gamma `$ is a numeric coefficient of order unity, and $`ϵ_0`$ plays the role of the high-energy cutoff (at higher energies the dispersion curve is no longer sound-like and physics is determined by single-particle processes). Equation (8) was first derived in Refs. . To find the decay rate $`\tau _I^1`$ one has to sum Eq. (8) over $`I^{}`$, but since we are interested in the parameter range $`\alpha 1`$ and $`Tϵ_0`$ the dominant contribution comes from $`I^{}=I\pm 1`$. It follows then from Eq. (3) that $`\xi 4\pi ^2n_sI/mLT`$ even for $`E_IT`$, and we may neglect current energy transfer to the bosonic environment in Eq. (8). Also, for $`\alpha 1`$ we may write approximately $`\mathrm{\Gamma }[1+\alpha ]/\mathrm{\Gamma }[1/2+\alpha ]\sqrt{\alpha }`$. The final result may be written then as $$\tau _I^1\frac{|gV_0n|^2}{T\sqrt{\alpha }}\left(\frac{2\pi T}{\gamma ϵ_0}\right)^{2\alpha }.$$ (12) We see that for large $`\alpha `$ the decay rate is severely suppressed at low temperatures $`T<T_0ϵ_0/2\pi `$. Physically, this effect reflects small overlap between different vacuum states corresponding to quantum numbers $`I`$ and $`I^{}`$. In practice, for $`\alpha 1`$ there is a “kinetic crossover” between the superfluid and normal behavior, i.e., between frozen and fast current relaxation, at $`TT_0`$. To estimate realistic parameters for ultra-cold atomic gases we have to start from the original quasi-1D geometry. Let the transverse motion is confined by the symmetric parabolic potential with frequency $`\omega _{}`$. Assuming that the 3D scattering length $`a`$ is small as compared with the oscillator length $`l_{}=1/\sqrt{m\omega _{}}`$, we may derive the potential energy per particle from the 3D relation $$ϵ_0=n\frac{4\pi a}{m}𝑑𝐫_{}\phi _0^4(r_{})nU_{\mathrm{eff}},$$ (13) where $`\phi _0`$ is the wavefunction of the transverse motion (here $`n`$ is the 1D particle density). For the parabolic potential we have $$ϵ_0=n2a\omega _{}.$$ (14) Purely 1D kinematics requires two conditions to be satisfied $$T\omega _{},ϵ_0\omega _{}.$$ (15) The second condition requires that a specific gas parameter, $`na`$, be small: $`na1`$. When deriving Eq. (12) we assumed that $`Tϵ_0`$ to guarantee that transition rates between the supercurrent states are negligible. Obviously, this is a more severe restriction on the temperature range than the first inequality in Eq. (15). We now turn to the parameter $`K`$ for a translationally invariant system. (Weak disorder does not change the value of $`K`$ dramatically.) For the weakly interacting gas compressibility is given simply by inverse $`U_{\mathrm{eff}}`$, and, in the low-energy limit, $`n_s=n`$. Thus, $$K=\frac{1}{\pi }\sqrt{\frac{mU_{\mathrm{eff}}}{n}}=\frac{\sqrt{2}}{\pi }\left(\frac{a}{l_{}}\right)^{1/2}\frac{1}{(nl_{})^{1/2}}.$$ (16) Experimentally it causes no difficulty to choose $`nl_{}>1`$, and to make index $`K`$ very small or parameter $`\alpha =1/K1`$. Let us consider an example of sodium gas assuming 1D density $`n=10^6`$ cm<sup>-1</sup>. Then, $`na0.25`$ and $`ϵ_00.5\times \omega _{}`$. For the magnetic trap frequency $`\omega _{}=3\times 10^4`$ s<sup>-1</sup>, we find $`nl_{}30`$, and extremely small $`K<0.01`$ (!). The crossover temperature for the frozen current dynamics is roughly $`T_0=ϵ_0/2\pi 2.5\times 10^8`$ K, and within the experimental range. We note that the effect of supercurrent stability considered above is essentially collective in nature and non-perturbative. Our transition Hamiltonian is valid only if initial and final states are given in terms of sound-like bosonic excitations - its validity may be questioned if the energy change in the transition, $`\xi `$, exceeds $`ϵ_0`$. This imposes the restriction that interaction is not macroscopically small, $`U_{\mathrm{eff}}2\pi ^2/mL`$. Also, even formally, for $`\xi >ϵ_0`$ we find very fast relaxation rates from Eq. (8) (for estimates one may roughly replace in Eq. (12) $`2\pi T\xi `$). We would like thus to verify that with all the above parameters we still may neglect energy (and momentum) transfer between the current and bosonic modes. Since $`\xi =E_IE_{I1}=4\pi ^2n(I1/2)/mL`$, for $`L=1`$ cm we find $`\xi 8\times 10^9(I1/2)`$ K, that is in the temperature range between $`4\times 10^9`$ K and $`T_0`$ we still have $`\xi <T`$ for thermal values of $`I`$ and system-size dependence is not crucial. On another hand, even if the energy has been dissipated into a single bosonic mode (in fact roughly $`\alpha 1`$ bosons are emitted/absorbed in the transition), the momentum of the boson would be only $`q=4\pi (I1/2)/KL`$ and much smaller than the momentum change of the current state $`2\pi n`$. Thus, a quasi-1D interacting Bose gas with realistic parameters can support a supercurrent at $`T0`$. The stability of such a state is guaranteed by extremely large relaxational time as compared to any reasonable experimental time. In connection with the question of preparation of supercurrent state, it is worthwhile to note that for an equilibrium 1D system at finite temperature there is a finite probability to be found in a supercurrent state. This naturally follows from the fact that the energy of a supercurrent state scales like $`1/L`$, so that at finite $`T`$ the statistical ensemble involves a large set of different numbers $`I`$. After a deep cooling, the system will be found in a state with some particular $`I`$, randomly varying from one experiment to another . This work was supported by the Russian Foundation for Basic Research (Grant No. 98-02-16262) and by the Grant INTAS-97-0972 \[of the European Community\].
no-problem/9908/gr-qc9908057.html
ar5iv
text
# Untitled Document Causality and Locality in Modern Physics, 171–178 G. Hunter, S. Jeffers, J.-P. Vigier (eds.) $`\mathrm{\copyright }`$ 1998 Kluwer Academic Publishers THE ZERO-POINT FIELD AND INERTIA B. HAISCH Solar and Astrophysics Laboratory, Lockheed Martin 3251 Hanover St., Palo Alto, CA 94304 A. RUEDA Dept. of Electrical Engineering & Dept. of Physics California State Univ., Long Beach, CA 90840 1. Introduction Is the vacuum electromagnetic zero-point field (ZPF) real? In a nice collection of examples, Milonni (1988) has shown that the interpretations of vacuum field fluctuations vs. radiation reaction are merely like two sides of the same quantum mechanical coin (cf. Senitzky 1973). Physical phenomena such as spontaneous emission, the Lamb shift and the Casimir force can be analyzed either way with the same result. The Casimir force is of particular interest (Milonni 1982). The recent measurements by Lamoreaux (1997) show agreement with the semiclassical theory of Casimir based on a real ZPF to within 5% over the measured range. Of course this effect is derived in standard QED calculations via subtraction of two formally infinite integrals over electromagnetic field modes. Another approach yielding identical results simply treats the quantum vacuum as consisting of (virtual) photons carrying linear momentum. Reflections off the conducting plates inside and outside the cavity are in balance for wavelengths shorter than the plate separation, but for longer wavelengths modes are excluded within the cavity. This imbalance results in a net “zero-point radiation pressure” pushing the plates together which is exactly the Casimir force (Milonni, Cook & Goggin 1988; Milonni 1994). One may argue over the correct theoretical perspective, but the new measurements leave no doubt that the predicted macroscopic forces are quite real. A similar treatment of the quantum vacuum can be applied to scattering of momentum-carrying zero-point photons by quarks and electrons in matter. As with the Casimir cavity, such scattering is almost entirely a detailed balance process. However it can be shown that, owing to acceleration effects within the class of those first studied by Davies (1975) and Unruh (1976), an acceleration-dependent imbalance results in a net reaction force (Rueda & Haisch 1997, 1998). Thus as with the Casimir force, a zero-point field quantum vacuum effect is proposed to give rise to a macroscopic phenomenon: in this case, the inertia of matter. 2. The Zero-Point Field in Quantum Physics The Hamiltonian of a one-dimensional harmonic oscillator of unit mass may be written (cf. Loudon 1983, chap. 4) $$\widehat{H}=\frac{1}{2}(\widehat{p}^2+\omega ^2\widehat{q}^2),$$ $`(1)`$ where $`\widehat{p}`$ is the momentum operator and $`\widehat{q}`$ the position operator. From these the destruction (or lowering) and creation (or raising) operators are formed: $$\widehat{a}=(2\mathrm{}\omega )^{1/2}(\omega \widehat{q}+i\widehat{p}),$$ $`(2a)`$ $$\widehat{a}^{\mathrm{}}=(2\mathrm{}\omega )^{1/2}(\omega \widehat{q}i\widehat{p}).$$ $`(2b)`$ The application of these operators to states of a quantum oscillator results in lowering or raising of the state: $$\widehat{a}|n=n^{1/2}|n1,$$ $`(3a)`$ $$\widehat{a}^{\mathrm{}}|n=(n+1)^{1/2}|n+1.$$ $`(3b)`$ Since the lowering operator produces zero when acting upon the ground state, $$\widehat{a}|0=0,$$ $`(4)`$ the ground state energy of the quantum oscillator, $`|0`$, must be greater than zero, $$\widehat{H}|0=E_0|0=\frac{1}{2}\mathrm{}\omega |0,$$ $`(5)`$ and thus for excited states $$E_n=\left(n+\frac{1}{2}\right)\mathrm{}\omega .$$ $`(6)`$ The electromagnetic field is quantized by associating a quantum mechanical harmonic oscillator with each k-mode. Plane electromagnetic waves propagating in a direction k may be written in terms of a vector potential $`𝐀_𝐤`$ as (ignoring polarization for simplicity) $$𝐄_𝐤=i\omega _𝐤\{𝐀_𝐤\mathrm{exp}(i\omega _𝐤t+i𝐤𝐫)𝐀_𝐤^{}\mathrm{exp}(i\omega _𝐤ti𝐤𝐫)\},$$ $`(7a)`$ $$𝐁_𝐤=i𝐤\times \{𝐀_𝐤\mathrm{exp}(i\omega _𝐤t+i𝐤𝐫)𝐀_𝐤^{}\mathrm{exp}(i\omega _𝐤ti𝐤𝐫)\}.$$ $`(7b)`$ Using generalized mode coordinates analogous to momentum ($`P_𝐤`$) and position ($`Q_𝐤`$) in the manner of (2ab) above one can write $`𝐀_𝐤`$ and $`𝐀_𝐤^{}`$ as $$𝐀_𝐤=(4ϵ_0V\omega _𝐤^2)^{\frac{1}{2}}(\omega _𝐤Q_𝐤+iP_𝐤)\epsilon _𝐤,$$ $`(8a)`$ $$𝐀_𝐤^{}=(4ϵ_0V\omega _𝐤^2)^{\frac{1}{2}}(\omega _𝐤Q_𝐤iP_𝐤)\epsilon _𝐤.$$ $`(8b)`$ In terms of these variables, the single-mode energy is $$<E_𝐤>=\frac{1}{2}(P_𝐤^2+\omega _𝐤^2Q_𝐤^2).$$ $`(9)`$ Equation (8) is analogous to (2), as is Equation (9) with (1). Just as mechanical quantization is done by replacing $`𝐱`$ and $`𝐩`$ by quantum operators $`\widehat{𝐱}`$ and $`\widehat{𝐩}`$, so is the quantization of the electromagnetic field accomplished by replacing $`𝐀`$ with the quantum operator $`\widehat{𝐀}`$, which in turn converts $`𝐄`$ into the operator $`\widehat{𝐄}`$, and $`𝐁`$ into $`\widehat{𝐁}`$. In this way, the electromagnetic field is quantized by associating each k-mode (frequency, direction and polarization) with a quantum-mechanical harmonic oscillator. The ground-state of the quantized field has the energy $$<E_{𝐤,0}>=\frac{1}{2}(P_{𝐤,0}^2+\omega _𝐤^2Q_{𝐤,0})^2=\frac{1}{2}\mathrm{}\omega _𝐤.$$ $`(10)`$ 3. The Zero-Point Field in Stochastic Electrodynamics Stochastic Electrodynamics (SED; see de la Peña & Cetto 1996; Milonni 1994) treats the ZPF via a plane electromagnetic wave modes expansion representation whose amplitudes are exactly such as to result in a phase-averaged energy of $`\mathrm{}\omega /2`$ in each mode (k,$`\sigma `$), where $`\sigma `$ represents polarization (cf. Boyer 1975): $$𝐄^{ZP}(𝐫,t)=\mathrm{Re}\underset{\sigma =1}{\overset{2}{}}d^3k\widehat{\epsilon }_{𝐤,\sigma }\left[\frac{\mathrm{}\omega _𝐤}{8\pi ^3ϵ_0}\right]^{\frac{1}{2}}\mathrm{exp}(i𝐤𝐫i\omega _𝐤t+i\theta _{𝐤,\sigma }),$$ $`(11a)`$ $$𝐁^{ZP}(𝐫,t)=\mathrm{Re}\underset{\sigma =1}{\overset{2}{}}d^3k(\widehat{k}\times \widehat{\epsilon }_{𝐤,\sigma })\left[\frac{\mathrm{}\omega _𝐤}{8\pi ^3ϵ_0}\right]^{\frac{1}{2}}\mathrm{exp}(i𝐤𝐫i\omega _𝐤t+i\theta _{𝐤,\sigma }).$$ $`(11b)`$ This kind of representation was used by Planck (1914) and Einstein and co-workers (Bergia 1979). The stochasticity is entirely in the phase, $`\theta _{𝐤,\sigma }`$, of each wave. (As discussed in $`\mathrm{\S }`$ 7 this is not entirely correct.) The spectral energy density of the classical ZPF is obtained from the number of modes per unit volume, $`8\pi \nu ^2/c^3`$ (Loudon 1983, Eq. 1.10), times the energy per mode, $`h\nu /2`$. The Planck spectrum plus ZPF radiation is thus: $$\rho (\nu ,T)d\nu =\frac{8\pi \nu ^2}{c^3}\left(\frac{h\nu }{e^{h\nu /kT}1}+\frac{h\nu }{2}\right)d\nu .$$ $`(12)`$ 4. The Davies-Unruh Effect Motivated by Hawking’s evaporating black hole concept, Davies (1975) and Unruh (1976) determined that a Planck-like component of the background scalar field will arise as seen from a uniformly-accelerated point with constant proper acceleration a (where $`|𝐚|=a`$) having an effective temperature, $$T_a=\frac{\mathrm{}a}{2\pi ck}.$$ $`(13)`$ (For the classical Bohr electron, $`v^2/r10^{25}`$ cm/s<sup>2</sup>, $`T_a367`$ K.) This effect is derivable from quantum field theory (Davies 1975, Unruh 1976). It was also derived in SED for the classical ZPF by Boyer (1980) who obtained for the spectrum a quasi-Planckian form (in the absence of external radiation): $$\rho (\nu ,T_a)d\nu =\frac{8\pi \nu ^2}{c^3}\left[1+\left(\frac{a}{2\pi c\nu }\right)^2\right]\left[\frac{h\nu }{2}+\frac{h\nu }{e^{h\nu /kT_a}1}\right]d\nu .$$ $`(14)`$ 5. Newtonian Inertia from ZPF Electrodynamics While these additional acceleration-dependent terms in Eq. (14) do not show any spatial asymmetry in the expression for the ZPF spectral energy density, certain asymmetries do appear when the electromagnetic field interactions with charged particles are analyzed. Haisch, Rueda & Puthoff (HRP; 1994) made use of this to propose a connection between the ZPF and inertia of matter. Assume that there are interactions between a real ZPF, represented as above, and matter at the fundamental particle level, treated as a collection of electrons and quarks, both of which are simply thought of as oscillating point charges: partons in the terminology of Feynmann. If the ZPF-parton interactions take place at high frequencies, then one need not worry about how the three quarks in a proton or a neutron are bound together. Each will interact independently with the ZPF, even though the three-quark ensemble is constrained to macroscopically move together. The method of Einstein and Hopf (1910) was followed: it breaks the analysis of the dynamics of the uniformly-accelerated parton into two steps. First we assume that the electric component of the ZPF, $`𝐄^{zp}`$, drives the parton to harmonic oscillation, i.e. creates a Planck oscillator. For simplicity we restrict these oscillations to a $`yz`$-plane characterized by the velocity vector $`𝐯_{osc}`$ and we force the oscillating parton to accelerate, via an external agent, in the $`x`$-direction with constant acceleration $`𝐚`$ (perpendicular to $`𝐯_{osc}`$). The acceleration will introduce asymmetries in the ZPF radiation field perceived by the oscillating parton. Second, we then ask what the effect is of the magnetic ZPF-parton interactions, specifically, what is the resulting Lorentz force: $`<𝐯_{osc}\times 𝐁^{zp}>`$? The result was the discovery of a reaction force of the form $$𝐅_r=𝐯_{osc}\times 𝐁^{zp}=\left[\frac{\mathrm{\Gamma }_Z\mathrm{}\omega _c^2}{2\pi c^2}\right]𝐚.$$ $`(15)`$ The quantity in brackets on the right hand side we interpreted as the inertial mass, $$m_i=\left[\frac{\mathrm{\Gamma }_Z\mathrm{}\omega _c^2}{2\pi c^2}\right]$$ $`(16)`$ where $`\mathrm{\Gamma }_Z`$ is the classical radiation damping constant of Abraham and Lorentz, but now referring to the Zitterbewegung oscillations <sup>c</sup> As discussed in chapter 17 of Jackson (1975) Classical Electrodynamics, one can obtain a characteristic radiation damping time for an electron having the value $`\mathrm{\Gamma }_e=6.26\times 10^{24}`$. This is not the proper $`\mathrm{\Gamma }_Z`$ for Zitterbewegung. and $`\omega _c`$ was taken to be an effective cut-off frequency of either the ZPF spectrum itself (perhaps at the Planck frequency) or of the particle-field interaction owing to a minimum (Planck) particle size (Rueda 1981). Newton’s Third Law tells us that a (motive) force, $`𝐅`$ will generate an equal and opposite (reaction) force, $`𝐅_r`$, and from Equation (15) $$𝐅=𝐅_r=m_i𝐚.$$ $`(17)`$ Newton’s third law is fundamental, whereas Newton’s second law, $`𝐅=m𝐚`$, appears to be derivable from the third law together with the laws of electrodynamics. 6. The Relativistic Formulation of ZPF-based Inertia The oversimplification of an idealized oscillator interacting with the ZPF as well as the mathematical complexity of the HRP analysis are understandable sources of skepticism, as is the limitation to Newtonian mechanics. A relativistic form of the equation of motion having standard covariant properties has been obtained (Rueda & Haisch 1998), which is independent of any particle model, relying solely on the standard Lorentz-transformation properties of the electromagnetic fields. Newton’s third law states that if an agent applies a force to a point on an object, at that point there arises an equal and opposite force back upon the agent. Were this not the case, the agent would not experience the process of exerting a force and we would have no basis for mechanics. The mechanical law of equal and opposite contact forces is thus fundamental both conceptually and perceptually, but it is legitimate to seek further underlying connections. In the case of a stationary object (fixed to the earth, say), the equal and opposite force can be said to arise in interatomic forces in the neighborhood of the point of contact which act to resist compression. This can be traced more deeply still to electromagnetic interactions involving orbital electrons of adjacent atoms or molecules, etc. A similar experience of equal and opposite forces arises in the process of accelerating (pushing on) an object that is free to move. It is an experimental fact that to accelerate an object a force must be applied by an agent and that the agent will thus experience an equal and opposite reaction force so long as the acceleration continues. It appears that this equal and opposite reaction force also has a deeper physical cause, which turns out to also be electromagnetic and is specifically due to the scattering of ZPF radiation. Rueda & Haisch (1998) demonstrate that from the point of view of the pushing agent there exists a net flux (Poynting vector) of ZPF radiation transiting the accelerating object in a direction opposite to the acceleration. The scattering opacity of the object to the transiting flux creates the back reaction force called inertia. The new approach is less complex and model-dependent than the HRP analysis in that it assumes simply that the fundamental particles in any material object interact with the ZPF in some way that is analogous to ordinary scattering of radiation. It is well known that treating the ZPF-particle interaction as dipole scattering is a successful representation in that the dipole-scattered field exactly reproduces the original unscattered field radiation pattern, i.e. results in detailed balance. It is thus likely that dipole scattering is a correct way to describe the ZPF-particle interaction, but in fact for our analysis we simply need to assume that there is some dimensionless efficiency factor, $`\eta (\omega )`$, that describes whatever the process is (be it dipole scattering or not). We suspect that $`\eta (\omega )`$ contains one or more resonances, but again this is not a necessary assumption. The new approach relies on making standard transformations of the $`𝐄^{zp}`$ and $`𝐁^{zp}`$ from a stationary to an accelerated coordinate system (cf. $`\mathrm{\S }`$ 11.10 of Jackson, 1975). In a stationary or uniformly-moving frame the $`𝐄^{zp}`$ and $`𝐁^{zp}`$ constitute an isotropic radiation pattern. In an accelerated frame the radiation pattern acquires asymmetries. There is thus a non-zero Poynting vector in any accelerated frame carrying a non-zero net flux of electromagnetic momentum. The scattering of this momentum flux generates a reaction force, $`𝐅_r`$. Moreover since any physical object will undergo a Lorentz contraction in the direction of motion the reaction force, $`𝐅_r`$, can be shown to depend on $`\gamma _\tau `$, the Lorentz factor (which is a function of proper time, $`\tau `$, since the object is accelerating). We find that $$m_i=\left[\frac{V_0}{c^2}\eta (\omega )\frac{\mathrm{}\omega ^3}{2\pi ^2c^3}𝑑\omega \right].$$ $`(18)`$ We find the momentum of the object to be of the form $$𝐩=m_i\gamma _\tau 𝐯_\tau .$$ $`(19)`$ Thus, we arrive at the relativistic equation of motion $$=\frac{d𝒫}{d\tau }=\frac{d}{d\tau }(\gamma _\tau m_ic,𝐩).$$ $`(20)`$ The origin of inertia becomes remarkably intuitive. Any material object resists acceleration because the acceleration produces a perceived flux of radiation in the opposite direction that scatters within the object and thereby pushes against the accelerating agent. Inertia is a kind of acceleration-dependent electromagnetic drag force acting upon fundamental charges particles. 7. For the Future Clearly a quantum field theoretical derivation of the ZPF-inertia connection is highly desireable. Another approach would be to demonstrate the exact equivalence of SED and QED. However as shown convincingly by de la Peña and Cetto (1996), the present form of SED is not compatible with QED, but modified forms could well be, such as their own proposed “linear SED.” Another step in the direction of reconciling SED and QED is the proposed modification of SED by Ibison and Haisch (1996), who showed that a modification of the standard ZPF representation (Eqs. 11ab) can exactly reproduce the statistics of the electromagnetic vacuum of QED. For an oscillator of amplitude $`\pm A`$ the classical probability of finding the point-mass in the interval $`dx`$ is a smooth function with a minimum at the origin (where the velocity is greatest) and a maximum at the endpoints of the oscillation. Treated quantum mechanically, an oscillator has a very different behaviour, but in an excited state approximates the classical probability distribution in the mean (see Fig. 1 of Ibison & Haisch). However the quantum $`n=0`$ ground-state — the one of direct relevance to the ZPF — is radically different from the classical one: the quantum-state probability maximum occurs where the classical state probability is at a minimum (position zero) and vice versa at the endpoints; indeed the quantum probability distribution is non-zero beyond $`\pm A`$. In both cases the average position, of course, remains zero. The same disagreement characterizes the difference between the Boyer description of the ZPF and the quantum ZPF. It has been shown by Ibison and Haisch (1996) that this can be remedied by introduction of a stochastic element into the amplitude of each mode that precisely agrees with the quantum statistics. This gives us confidence that the SED basis of the inertia (and gravitation) concepts is a valid one. The two most frequently posed questions, indeed perhaps the most important ones, are (1) whether the ZPF-inertia theory is subject to experimental validation, and (2) what the implications might be for revolutionary new technologies. An independent assessment of the case for experimental testing was carried out by Forward (1996) as a USAF-sponsored study. No direct test could be identified as currently feasible, but a constellation of related experiments were identified. A NASA Breakthrough Propulsion Physics program is being initiated, and the ZPF-inertia concept is high on the list of candidate ideas to explore (see Haisch & Rueda 1997a) along with the Sakharov-Puthoff concept of ZPF-gravitation linked to the ideas herein by the principle of equivalence (Sakharov 1968, Puthoff 1989, see also Haisch & Rueda 1997b). We note that four decades elapsed before atomic energy became a technology following Einstein’s 1905 paper proposing (special) relativity. A similar time-scale may apply here. Acknowledgements We acknowledge support of NASA contract NASW-5050 for this work. BH also thanks Prof. J. Trümper and the Max-Planck-Institut where some of these ideas originated during several extended stays as a Visiting Fellow. AR acknowledges valuable discussions with Dr. D. C. Cole. References Bergia, S. (1979), The light quantum and the wave-particle duality for radiation in Einstein’s research: 1904–1925, preprint, Inst. di Fisica, U. Bologna. Boyer, T.H. (1975), Random electrodynamics: The theory of classical electrodynamics with classical electromagnetic zero-point radiation, Phys. Rev. D, 11, 790–808. Boyer, T.H. (1980), Thermal effects of acceleration through random classical radiation, Phys. Rev. D, 21, 2137–2148. Boyer, T.H. (1984), Thermal effects of acceleration for a classical dipole oscillator in classical electromagnetic zero-point radiation, Phys. Rev. D, 29, 1089–1095. Davies, P.C.W. (1975), Scalar particle production in Schwarzschild and Rindler metrics, J. Phys. A, 8, 609. de la Peña, L. & Cetto, A. (1996), The Quantum Dice: An Introduction to Stochastic Electrodynamics, (Kluwer Acad. Publ., Dordrecht). Einstein, A. & Hopf, L. (1910), Über einen Satz der Wahrscheinlichkeitsrechnung und seine Anwendung in der Strahlungstheorie, Annalen der Physik (Leipzig), 33, 1096; Statistische Untersuchung der Bewegung eines Resonators in einem Strahlungsfeld, 33, 1105. Forward, R.L. (1996), Mass Modification Experiment Definition Study, Phillips Lab. Rept. 96-3004; also published in J. Sci. Expl., 10, 325–354 (1996). Haisch, B. & Rueda, A. (1997a), The Zero-Point Field and the NASA Challenge to Create the Space Drive, Proc. NASA Breakthrough Prop. Physics Workshop, in press. Haisch, B. & Rueda, A. (1997b), Reply to Michel’s ‘Comment on Zero-Point Fluctuations and the Cosmological Constant’, Ap. J., 488, 563. Haisch, B., Rueda, A. & Puthoff, H.E. (1994), Inertia as a zero-point field Lorentz Force, Phys. Rev. A, 49, 678. Ibison, M. & Haisch, B. (1996), Quantum and classical statistics of the electromagnetic zero-point field, Phys. Rev. A, 54, 2737. Jackson, J. D. (1975), Classical Electrodynamics, 2nd ed., Wiley & Sons. Lamoreaux, S.K. (1997), Demonstration of the Casimir Force in the 0.6 to 6 $`\mu `$m Range, Phys. Rev. Letters 78, 5–8. Loudon, R. (1983), The Quantum Theory of Light (2nd. ed)., Oxford Univ. Press, chap. 4. Milonni, P.W. (1982), Casimir forces without the radiation field, Phys. Rev. A 25, 1315–1327. Milonni, P.W. (1988), Different Ways of Looking at the Electromagnetic Vacuum, Physica Scripta T21, 102–109.. Milonni, P.W. (1994), The Quantum Vacuum, Academic Pres, chap. 1. Milonni, P.W., Cook, R.J. & Goggin, M.E. (1988), Radiation pressure from the vacuum: Physical interpretation of the Casimir force, Phys. Rev. A 38, 1621–1623. Planck, M. (1914), The theory of heat radiation (P. Blakansten & Son, London) Puthoff, H.E. (1989), Gravity as a zero-point fluctuation force, Phys. Rev. A, 39, 2333. Rueda, A. (1981), Behaviour of classical particles immersed in the classical electromagnetic field, Phys. Rev. A, 23, 2020. Rueda, A. & Haisch, B. (1997), Electromagnetic vacuum and inertial mass, this volume. Rueda, A. & Haisch, B. (1998), Inertial mass as reaction of the vacuum to accelerated motion Phys. Lett. A, 240, 115. Sakharov, A. (1968), Vacuum Quantum Fluctuations in Curved Space and the Theory of Gravitation Soviet Physics - Doklady, 12, No. 11, 1040. Senitzky, I.R. (1973), Radiation reaction and vacuum field effects in Heisenberg picture QED, Phys. Rev. Lett., 31, 955. Unruh, W.G. (1976) Notes on black-hole evaporation, Phys. Rev. D, 14, 870.
no-problem/9908/cond-mat9908269.html
ar5iv
text
# Why Self-Consistent Diagrammatic Perturbation Theory is ”just” Perturbation Theory ## I The Arguments In this write-up, the advantages of the sea-boson technique are highlighted and contrasted with the short-comings of self-consistent diagrammatic perturbation theory(i.e. Feynman diagrams or its algebraic counterpart introduced by Schwinger). For more details the reader is refered to our published work . Here and henceforth the term ”the text” refers to the book by Kadanoff and Baym. For the sake of definiteness we shall focus here only on fermions and that too without spin(S=0) therefore we shall employ only one of the signs used in the text namely the lower sign corresponding to fermions. All equation numbers starting with the letters ”KB” are from this book. Let us start with the equation of motion for the Green function(KB 12-7). $$(i\frac{}{t_1}+\frac{_1^2}{2m}U_{eff}(1))G(1,1^{^{}};U)=\delta (11^{^{}})+d2\text{ }\mathrm{\Sigma }^{^{}}(1,2;U_{eff})G(2,1^{^{}};U_{eff})$$ (1) where, $$U_{eff}(1)=U(1)+d2\text{ }v(12)[i\text{ }G(2,2^+;U_{eff})n_0]$$ (2) here $`n_0`$ is the mean density of particles. $$\mathrm{\Sigma }^{^{}}(1,1^{^{}};U_{eff})=i\text{ }v_S(1,1^{^{}};U_{eff})G(1,1^{^{}};U_{eff})+id3\text{ }d4\text{ }v_S(1,3;U_{eff})G(1,4;U_{eff})\frac{\delta \mathrm{\Sigma }^{^{}}(4,1^{^{}};U_{eff})}{\delta U_{eff}(3)}$$ (3) The $`v(12)=\lambda \text{ }v(𝐱_1𝐱_2)\delta (t_1t_2)`$ is the bare Coulomb interaction and $`G(1,2;U=0)`$ is the full one-particle Green function in equilibrium and $`\lambda `$ is a dimensionless coupling constant. Further(KB 12-10), $`v_S(1,3)=v(13)i{\displaystyle d2\text{ }d4\text{ }v_S(1,2)G(4,2)G(2,4^+)v(43)}`$ $$iv_S(1,2)G(4,5)\frac{\delta \mathrm{\Sigma }^{^{}}(5,5^{^{}})}{\delta U_{eff}(2)}G(5^{^{}},4)v(43)$$ (4) As it stands the above sets of equations are defined independent of perturbation theory. That is, $`\lambda `$ is not assumed to be small. We wish to argue that within the framework of perturbation theory, the Green function obtained from the above equations has a simple analytic structure in the vicinity of the origin of the coupling constant space ($`\lambda =0`$ ). More precisely, we argue that, following the letter and spirit of perturbation theory leads us to the inevitable conclusion that the one-particle Green function is an analytic function of the coupling constant, perhaps with zero radius of convergence. This is true even if the Green function is obtained ”self-consistently” (that is, by solving a system of coupled non-linear integro-differential equations). This state of affairs should be contrasted with the sea-boson method where we found that at least in one-dimension, the one-paricle Green function has a nonanalytic dependence in the coupling at the origin of the coupling constant space. Thus this feature is completely erased from diagrammatic perturbation theory due to the latter’s inherent and as it happens, drastic assumptions. The letter and spirit of perturbation theory demands that we solve for the correlation self-energy by expanding in powers of the (screened or shielded) Coulomb interaction. When this is done, it may be seen quite easily that the correlation self-energy is an analytic function of the coupling constant at the origin of the the coupling constant space. To see this more clearly let us first observe that the solutions to Eqs.( 1) and ( 3) together with KMS boundary conditions are unique. Therefore the Green functions and self-energies possess unique analytic structures. Furthermore, if an analysis shows that a Green function is (and self-energy) analytic at the origin of the coupling constant space is consistent with the above sets of equations then indeed this is the only possible analytic structure possessed by these quantities. This is what we shall now argue. Let us retain only the GW-part of the correlation self-energy, that is, neglect all derivatives of the self-energy with respect to the effective potential as suggested by Kadanoff and Baym. When this is done one may proceed to convince oneself of the analyticity of the Green function as follows. Since we are interested in the neighbourhood of $`\lambda =0`$, one can argue that the zeroth order approximation to the full Green function is $`G_0(1,1^{^{}})`$, namely the noninteracting one. Then one may use this to evaluate the self-energy and the screened Coulomb interaction. Note that so long as we are sufficiently close to $`\lambda =0`$, no loss of generality is entailed by this procedure. In particular, a ”self-consistent” solution to this system for $`\lambda `$ sufficently close to $`\lambda =0`$ is equivalent to the usual pertubative procedure we just indicated. To proceed with the description of the iterative procedure, then one may re-evaluate the full Green function by including linear terms in $`\lambda `$ on the right-hand side of Eq.( 1). This procedure when repeated leads us to the unambiguous conclusion that the full Green function is indeed an analytic function of $`\lambda `$ at $`\lambda =0`$. Two points are worth stressing again. First is that this conclusion depends rather strongly on the fact that we have chosen to ignore the derivatives of the self-energy with respect to the effective potential. Only then are we able to reduce the system to a familiar set of integro-differential equations rather then a set of functional equations. The second is that having convinced ourselves that a Green function that is analytic at $`\lambda =0`$ is consistent with the above sets of equations, we are also led to the conclusion that this structure is unique since the Green function itself is unique. Thus if a skeptical reader tries to argue that maybe if one had started with an ansatz for the full Green function that is non-analytic in $`\lambda `$ then one would have found that this is also consistent with the above sets of equations. Not so. The reason as we just alluded to is uniqueness. Having said this, we would now like to highlight some of the advantages of the sea-boson method that we have introduced into the literature. There we found that for small enough values of the coupling, the momentum distribution(or the equal-time component of the full Green function) is nonanalytic in the coupling, thus demonstrating that the sea-boson method is superior. For the sake of completeness, we reproduce the results here. $`c_k^{}c_k=n_F(k)+(2\pi k_F){\displaystyle _{\mathrm{}}^+\mathrm{}}\text{ }{\displaystyle \frac{dq_1}{2\pi }}\text{ }{\displaystyle \frac{\mathrm{\Lambda }_{kq_1/2}(q_1)}{2\omega _R(q_1)(\omega _R(q_1)+\omega _{kq_1/2}(q_1))^2(\frac{m^3}{q_1^4})(cosh(\lambda (q_1))1)}}`$ $$(2\pi k_F)_{\mathrm{}}^+\mathrm{}\text{ }\frac{dq_1}{2\pi }\text{ }\frac{\mathrm{\Lambda }_{k+q_1/2}(q_1)}{2\omega _R(q_1)(\omega _R(q_1)+\omega _{k+q_1/2}(q_1))^2(\frac{m^3}{q_1^4})(cosh(\lambda (q_1))1)}$$ (5) here, $$\lambda (q)=(\frac{2\pi q}{m})(\frac{1}{v_q})$$ (6) $$\omega _R(q)=(\frac{|q|}{m})\sqrt{\frac{(k_F+q/2)^2(k_Fq/2)^2exp(\lambda (q))}{1exp(\lambda (q))}}$$ (7) $$n_F(k)=\theta (k_F|k|),\text{ }\mathrm{\Lambda }_k(q)=n_F(k+q/2)(1n_F(kq/2))$$ (8) The interaction in the denominator present in the quantity $`\lambda (q)`$ means that the momentum distribution is nonanalytic in the coupling at $`v_q=0`$, thus making our claims concrete. The author wishes to thank Mr. Erich Mueller for critical comments that led to this write-up. Thanks are also due to Prof. J.J. Song and her group at Oaklahoma State University for their support both financially and morally. The author may be contacted at setlur@mrlxpa.mrl.uiuc.edu.
no-problem/9908/astro-ph9908242.html
ar5iv
text
# Chemical Evolution of Dwarf Irregular Galaxies – chemodynamical models and the effect of gas infall ## 1. Introduction ### 1.1. Morphologies and Star-formation Epochs of Dwarf Galaxies Dwarf galaxies (DGs) present a wide variety of morphological types. Their structural and chemical properties differ from those of giant galaxies. In addition, low-mass galaxies seem to form at all cosmological epochs and by different processes and from different sources. Dwarf elliptical galaxies (dEs) are an extremely common and astrophysically interesting class of galaxy. Most known dEs are found in regions with high galaxy densities, and they are the most numerous of all galaxy types in the cores of nearby galaxy clusters (see e.g., Wirth & Gallagher 1984, Binggeli 1994). Since the bulk of their star formation (SF) occurred in the past (see review by Ferguson & Binggeli 1994), dEs thus are frequently considered to be “stellar fossil” systems. The compact ”blue dropouts” (Madau et al. 1996) visible in the Hubble Deep Field (HDF; Williams et al. 1996) seem to reflect the formation epoch of dEs by part. Their number excess compared to the present dE density requires, however, the disappearance of some fraction, caused by three different processes: firstly, disruption by means of vehement evolutionary effects like starbursts (SBs) and subsequent intense type II supernova explosions (SNII) or, secondly, by means of tidal forces and last but not least due to their accretion by massive galaxies. In galaxies where gas is consumed by astration stellar abundances are predicted to be close to the solar value. On the other hand, the moderate-to-low stellar metallicities in dE and the related dwarf spheroidals (dSphs) which represent the low-mass end of DGs (about 0.1 of solar or less) suggest that extensive gas loss occurred during their evolution by means of SNII driven galactic winds (Larson 1974, Dekel & Silk 1986). Yet many dSph galaxies show not only a significant intermediate-age stellar population (Hodge 1989, Grebel 1997), but also more recent SF events (Smecker et al. 1994, Han et al. 1997) with increasing metallicity, indicating that gas was partly kept in the system. On the other hand, gas infall might also cause the excitation of a new SF episode like in NGC 205 (Welch et al. 1998). At medium redshifts faint blue galaxies appear (Colless et al. 1990, Lilly et al. 1991) which resemble spectroscopically local compact Hii regions with respect to their strong but narrow emission lines, their colors and the M/L ratio (Koo et al. 1995, Guzman et al. 1998, Guzman, this conference). Those extremely high SF rates exist at present also in a special DG type, which are characterized by small sizes, blue centrally bright colours, Hii-type emission lines, high gas content and low metallicity: blue compact DGs or also called Hii galaxies. Early survey studies e.g. by Haro, Zwicky, and Markarian have drawn attention to this class of objects. In general, gas-rich dwarf irregular galaxies (dIrrs) show a large variety of SF rates from low to extraordinarily high values as in SBs but consist additionally of at least one intermediate-age to old underlying stellar population like in NGC 1569 (Heckman et al. 1995) or NGC 1705 (Meurer et al. 1992). The newly formed stellar associations are often very massive and compact and, therefore, called super star clusters (SCC) which are assumed to evolve to Globular Clusters by aging (Ho & Fillipenko 1996). The addressed questions why and by what physical process such enormous SF rates are triggered, that would often consume all the gas content within less than a billion years, is not yet known. In most SBDG large Hi reservoirs enveloping the luminous galactic body have been recently detected (NGC 4449: Hunter et al. 1998, I Zw 18: van Zee et al. 1998b, NGC 1705: Meurer et al. 1998) and can to some extent be interpreted to fall in and, by this, to feed the SB (e.g. He 2-10: Kobulnicky et al. 1995). On the other hand, because of their low binding energy SBDGs are characterized by superwinds (Marlowe et al. 1995) or large expanding X-ray plumes which are driven by SNeII. Their existence demonstrates the occurrence of metal loss by means of large-scale galactic winds and allows to study this cosmologically common phenomenon still today. Even more smoothly developing dIrrs have experienced several SF epochs, like the LMC (Pagel & Tautvaisiene 1998). Their irregular shape, however, stems from their present mainly patchy distribution of Hii regions like it is visible e.g. in some prominent representatives NGC 4449 and NGC 4214. Although dIrrs consist of the same and higher gas fraction as giant spiral galaxies (gSs), they appear with a wide range but lower metallicity $`Z`$ as gSs. Again this implies that the metal-enriched gas from SNeII was lost from the galaxy. Vice versa also infall of low-metallicity (even primordial) gas can allow for this signature, but reasonably both might have acted. Nevertheless, very metal-poor dIrrs like e.g. I Zw 18 (Searle & Sargent 1972), GR8 (Skillman et al. 1988), or SBS 0335-52 (Thuan et al. 1997) seem to form their first generation of stars today. In addition, a strange but perhaps common mode of DG formation exists in tidal tails of merging galaxies (Mirabel et al. 1992, Duc & Mirabel 1998). Although these units seem to be gravitationally bound, their survival of energetic events like accumulative SNeII or tidal friction is questionable. ### 1.2. Element Abundances and the Chemical Evolution of Dwarf Irregulars While the low metal content in DGs can be attributed to both metal-enhanced mass loss and/or low-metallicity gas infall, another puzzling fact needs explanation: Why do DGs, though with O abundances below 1/10 solar, show also low N/O ratios of up to 0.7 dex smaller than in gSs with a large scatter and no significant correlation with O/H (Pagel 1985; see fig.1)? It is persuasive that normal disk Hii regions and stars are located along the line of secondary nitrogen production (SNP), while the N/O ratios of DGs scatter around an average value of -1.5 independently of their oxygen abundance. This is supposed to represent the primary production process of N (PNP) at low metallicities. Most interestingly, van Zee et al. (1998a) have recently found that the abundances of the outermost Hii regions in spiral galaxies at large radii have not only low O abundances, but spread over the same regime between SNP and PNP tracks and overlap with DG’s values in the sense that they approach the primary line (fig.2). For a further insight and deeper understanding of stellar yields and galactic element enrichment at early galactic stages one can compare N/O vs. O measurements of extragalactic objects from different cosmological ages. Abundances for quasar emission lines reach to very high abundances of one order of magnitude above solar (Hamann & Ferland 1992) as an extrapolation of the SNP line, while QSO absorption-line systems (ALS) scatter again without any strong tendency except that damped-Ly$`\alpha `$ systems (DLAs) (Molaro et al. 1996, 1998, Pettini et al. 1995) are more widely distributed than Lyman-limit systems (LLS) (Köhler et al. 1996) (see fig.3). Past standard stellar evolution models yield the following separation: While C and N are attributed by planetary nebulae (PNe) from intermediate-mass stars (IMS) to the warm cloudy gas phase (CM) of the interstellar medium (ISM), O and Fe are dominantly released by means of SNII and SNIa explosions, respectively, to the hottest phase of the ISM, the intercloud medium (ICM) and, therefore, trace these energetic events and their stellar progenitors. This differentiation of N and O pollution to the gas phases has led several authors to propose various chemical models which could successfully reach the observed N/O-O regime of DGs. Matteucci & Tosi (1985), Marconi et al. (1994) and Kunth et al. (1995) allow for galactic winds with selective element depletion, which are driven by numerous SB episodes, in order to account for the necessary reduction of O. Similarly, Garnett (1990) and Pilyugin (1992) present models that follow the assumption of abundance self-enrichment within the observed Hii regions as it was concluded before from observations (Kunth & Sargent 1986). All these models differentiate the O and N release due to discontinous and separate events in accordance to their progenitor lifetimes, by this, leading to saw tooth-like evolutionary tracks in the log(N/O)-log(O/H) diagram (Pilyugin 1992). Parameters like wind efficiency, IMF, and initial gas fraction provide sufficiently good model approaches to the observations. It should be emphasized that three important ingredients enter mainly these referred chemical evolutionary models: 1) The yields are in principle based on almost the same stellar evolution calculations (N by Renzini & Voli (1981) and O by Woosley & Weaver (1986) or by Maeder (1992) and Maeder & Meynet (1989)). 2) All the investigations start already from N/O values (even at low O abundance) reached by the element enrichment of a normal stellar population with traditional yields. 3) Gas infall of metal-poor gas is not considered because it reduces only the oxygen abundance while leaving N/O almost constant. If N and O are produced by different progenitor stars and are separately released to the different phases of the ISM, their simultaneous existence in Hii regions representing the ionized CM, could only result from mixing processes between CM and ICM. In models where these processes are taken into account, the observed N/O ratio should, therefore, permit qualitative studies on both the mixing direction and efficiency. Additionally, its radial distribution in a DG provides an insight into dynamical effects of the ISM. The assumption of an individual self-enrichment of Hii regions during their ionization-caused observability (Pilyugin 1992) would plausibly involve both, abundance variations during their visibility epoch and, by this, differences between Hii regions within the same dIrr. In contrast, Kobulnicky et al. (1999) report the non-detection of any sizable O, N, and He anomalies from Hii regions in the vicinity of young star clusters in SBDGs with one exception, NGC 5253, what reveals a central N overabundance (Kobulnicky et al. 1997). In NGC 1569, that has recently formed two SSCs, Kobulnicky & Skillman (1997) find constant O and N/O values over a radial extent of more than 600 pc with a scatter of N/O by only 0.2 dex, while self-enrichment tracks reach nearly one dex in N/O (Pilyugin 1992) and dispersal distances of much less. The same appearance is found in I Zw 18 on a scale of 1 kpc (Izotov 1999). From a study of five dIrrs Kobulnicky & Skillman (1998) stated that the hypothesis of differential mass (and by this element) loss is inconclusive. ## 2. Chemodynamical Models Although many authors concentrate on the more spectacular SBDGs, not all observed dIrrs that fill in the same N/O-O regime have passed or experience at present strong SBs, but form stars also at an almost moderate and continuous rate. Empirical studies and theoretical investigations under galaxy conditions of low gravitational potentials have proven that their ISM should be balanced by counteracting processes like SN heating and cooling, turbulence and dissipation (Burkert & Hensler 1989). Similarly, the SF itself is self-regulated under various conditions due to stellar heating, gas cooling, evaporation, and condensation (see e.g. Franco & Cox 1983, Köppen et al. 1995, 1998). If the evolution of galaxies is sensitive to the energetical impact from different processes, an adequate treatment of the dynamics of stellar and gaseous components and of their mutual energetical and materialistic interactions is essential. This modeling of galactic evolution is properly treated by the chemodynamical (cd) prescription. Its formulation in 1d and 2d dynamics with the ”materialistic” and ”energetic” equations can be found for interested readers e.g. in Theis, Burkert, & Hensler (1992) and Samland, Hensler, & Theis (1997), respectively. The evolution of DGs can proceed in self-regulated ways both, globally by large-scale flows of unbound gas but also locally in the SF sites. cd models of non-rotating DG systems achieve SBs from the initial collapse and after mass-loss induced expansion in the recollapse phase of its bound gas fraction (Hensler et al. 1993, 1999). Particular models are evolving also with oscillatory SF episodes. The formed stellar populations are similar to dSphs as well as more massive BCDGs and reveal a central concentration of the recent SF region embedded in older elliptical stellar populations. External effects (Vilchez 1995) like extended dark matter (DM) halos, the intergalactic gas pressure, etc., could cause further morphological differences of DGs e.g. by regulating the outflow and allowing for a recollapse or even fuelling subsequent SF by infall of intergalactic clouds. For rotating dIrrs cd models are essential which account at least for a 2d rotationally symmetric structure. 3d calculations (in progress) have to make use of an adaptive-mesh refinement algorithm for numerical reasons in order to find a compromise between acceptable computer time and sufficient spatial resolution. In our models a gaseous protogalactic cloud starts with or without a DM halo (Burkert 1995) and with a Plummer-Kuzmin-type gas distribution. In the following we wish to discuss rotating 10$`{}_{}{}^{9}\mathrm{M}_{}^{}`$ DG models which commence with a 2 kpc baryonic density scalelength. Recent stellar evolution models by Woosley & Weaver (1995), which provide a SNP (and a small amount of PNP) also by HMS are taken into account. For consistency also the most recent models for IMS by van der Hoek & Groenewegen (1997) have been applied. The evolutionary model of two 10$`{}_{}{}^{9}\mathrm{M}_{}^{}`$ models (with and without a DM halo) are shown in fig.3. Here the abundances of N and O are determined as the mean value in bright SF regions of the models in order to achieve the best comparability with observations of Hii regions. This new deduction of abundances differs from the already published model results (Hensler & Rieschick 1998, 1999) in the sense that the former abundances have been calculated as the mass-weighted mean of the whole computational domain, i.e. even with the halo gas, and all components. Although these former results provided already a first hint that the abundance ratios seem to deviate from simple models, they fail to be representative for the visible part. The evolutionary states of the models in fig.3 at the upper points of the tracks are shown at ages of 3.8 Gyrs for the DM model and of 3.4 Gyrs for the other without DM, respectively. Our cd models reveal three major and significant discrepancies to former above-mentioned chemical models: 1) Like normal dIrrs and analytical studies show, our models do not evolve into a global SB, but only to local SF enhancements (Rieschick & Hensler 1999). This leads to the conclusion, that SBs might need a trigger mechanism as observed in most SBDGs (see section 1.1). 2) In contrast to the metal enrichment in closed-box models, here the evolutionary track in the N/O-O diagram does not circumvent the regime of the observed values for dIrrs along a horizontal track at almost solar N/O ratios, but passes from low N/O and O/H values in the diagonal direction between PNP and SNP trends through the regime of dIrr observations (fig.3). 3) From fig.4 it becomes visible that N mixes perfectly with O in the ICM due to evaporation of the CM in the vicinity of the SF and SNII explosion sites. This achieves a nearly constant abundance ratio widely spread over large distances within the DG. The evaporation/condensation parameter, that is defined as $`\beta =(\rho _{cond}\rho _{evap})/(\rho _{cond}+\rho _{evap})`$, reachs -1.0 in the innermost center, while condensation starts to dominate outward of the steeper density gradient (also recognizable as $`\beta `$ gradient) where $`\beta `$ can amount to +0.2. While this latter fact lends supports to the observed radial constancy of the N and O abundances in a BCDGs, by this, it conflicts with the self-pollution argument for the Hii region except in the very center. In addtion, it contradicts to the selective expulsion of O by means of galactic winds (Matteucci & Tosi 1985). Obviously, a substantial mixing of CM and ICM determines the chemical evolution. To summarize, the dIrr abundance problem cannot be solved by SB-driven galactic wind, because the N/O ratio should remain constant. ## 3. Gas Infall Generally speaking and as a first still careful interpretation of the presented chemical evolution of cd dIrr models, one has to study the reasons for the occuring discrepancies of N/O abundance ratios between simple models and the cd models which, on the other hand, agree with DG observations. Although even with the new additional nitrogen production from HMS the shortly living IMS of around 5-8 $`\mathrm{M}_{}`$ should release their primary N on sufficiently short timescales and rise N/O to solar values. As a first step one has to compare the dynamical ”cd ” models with closed-box results under the inclusion of the same interaction prescriptions. Although the evaporation/condensation behaviour of the closed-box evolution might differ from the dynamical model and, by this, the element abundances in the gas phases should deviate, the global box model, i.e. the sum of all gaseous components, should gain the same abundance $`Z`$ in accordance to the analytical solution. Without going into too much detail here we wish to emphasize this agreement with the expected analytical behaviour and refer to a forthcoming paper (Rieschick & Hensler 1999). We conclude that this issue excludes the possible explanation that the discrepancy stems solely from the application of new stellar yields. More preferably, it must be caused by the dynamical behaviour of gas in a low gravitational potential with a high energy impact. Only if the SF rate is increasing steeply, the SNP of HMS can exceed the delayed release of primary N by IMS. Recently, Köppen & Edmunds (1999) have performed analytical studies of the effect of gas infall into a star-forming galaxy. Since the infalling gas is assumed to be primordial, it reduces the normal stellar yields in the sense that it decelerates the temporal metal enrichment or it approachs an equilibrium value. Formally, a constant ratio a=A/$`\mathrm{\Psi }`$ of temporal gas infall rate A(t) normalized to the star-formation rate $`\mathrm{\Psi }`$(t) yields a steadily increasing evolutionary path in the diagram joining purely secondary-to-primary element abundances $`Z_s/Z_p`$ vs. primary element. Also monotonically increasing or decreasing a(t) evolve along the secondary-production track, while for da/dt$`>`$0 an equilibrium point in this diagram will be reached at lower value than for constant or decreasing a(t). In a situation where significant gas infall occurs in a stochastic event, the enrichment path in the diagram makes a loop to lower $`Z_p`$ but turns back to the former track at lower $`Z_s/Z_p`$ when a(t) drops again to a state as before the infall episode. Even the beginning from an arbitrary point in this diagram cannot prevent to converge to the secondary production track again. While the formalism is described in their paper (Köppen & Edmunds 1999), the attention of interested readers should be drawn to an appropriate and instructive applet by one of us (Köppen 1998), where the analytical solutions of the set of equations can be comfortably visualized for different situations and different possible combination of variables. In fig.5 we wish to demonstrate how the observed scattered values of N/O with oxygen in dIrrs could be interpreted: After an initial infall epoch which represents the early collapse phase of the gaseous DG, a later episodic infall event can deflect the evolutionary track for the ratio of a purely secondarily produced element to a primary one towards lower single abundance at the same ratio. As can be discerned, the loop can pass through the regime of observed dIrr measurements. Our models can start from a basis of SNP with proceedingly enhanced PNP but additional gas infall episodes. The slope of the evolutionary path is then supposed to wobble between secondary and horizontal slope and because of small but varying gas infall rates. The chemical evolution of the cd models is directly and self-consistently passing through the regime of abundances observed in dIrrs (see fig.3). ## 4. Conclusions In this paper we have presented recent cd models of dIrrs that are aimed to investigate the reason of their observed peculiar element abundances and, by this, to address the question on their chemical evolution. Because of their capability to treat gas and stars dynamically within the framework of their interactions the self-consistent cd models discovered strong constraints on the assumptions of analytical considerations. In particular, mixing effects of the gas phases prevent a selective element depletion, and furthermore, the models evolve with local variations but globally moderate SF activity (Rieschick & Hensler 1999) and therefore typically for the major fraction of dIrrs. For actuality, the most appropriate yields from recent stellar evolution models for HMS (Woosley & Weaver 1995) and IMS (van der Hoek & Groenewegen 1997) are implied, where nitrogen is secondarily and (to less extent) primarily produced by HMS. Nevertheless, since the PNP in IMS raises reasonably the N/O ratio within a typical timescale of their stellar lifetimes to almost solar, i.e. larger than observed in dIrrs, a rapid O enhancement could be required. This invokes a present SB event as a plausible explanation. As a perspective, the observed low N/O ratios in DGs’ Hii regions, however, could be attained, if N would be purely produced as a secondary element in HMS. Nevertheless it is unavoidable, that the PNP by IMS should also dominate in a single stellar population after around 10<sup>8</sup> yrs and therefore enhances the N/O ratio. In contrast to the mentioned expectations, the chemical evolutionary track of the here presented cd model is surprising as it passes through the regime of observed abundances in dIrrs without any artificial manipulation. Three mechanisms could be considered for an explanation of this issue. Firstly, N and O are almost perfectly mixed in the innermost regions of dIrrs due to overwhelming evaporation of clouds embedded in the hot ISM and, by this, not allowing for a selective element loss in dIrrs. Although only a weak galactic wind might be produced in self-regulated low-mass galaxies, i.e. without any SB the centrally peaked SF drives a large-scale radial redistribution of N/O according to the presently exploding HMS in a surrounding ISM which is not yet completely polluted by primary N from IMS because of their longer lifetimes. The lower N/O hot gas mixture condensates at remote regions where it is also not in accordance with the locally existing stellar population. This issue involves that dIrrs are systems still forming from their large surrounding gas reservoir or strongly affected by it. Secondly, the cd results can be interpreted with respect to further gas infall in a global model where the SNP in HMS exceeds the PNP in IMS. For the first billion years this could be the case for a steeply increasing infall rate where the rapidly released HMS yields exceed the delayed PNP. At least, analytical studies show that just this latter situation can be reached repeatedly if gas infall episodes occur. By such events of different strength and durations the whole regime of secondary-to-primary element ratios in dIrrs can be covered. Since recent deep observations of dIrrs and SBDGs have discovered an increasing existence of large gas envelops, i.e. of enormous gas reservoirs circling or accumulating around the visible body of the BCDGs, the relevance of infall episodes for the evolution of these low-mass galaxies is doubtless and serves as the most promising explanation for their observed abundances and their rejuvenation. Therefore, this process is essential for evolutionary models. Since the SF can be triggered subsequently by such gas infall, SBs are a plausible consequence. Although the smooth collapse of a protogalactic cloud like in our conditions resembles somehow a steady, but decreasing gas infall rate, stochastic infall of intergalactic gas clouds is not yet included in the presented models. As a preliminary result from this discussion one should, however, conclude that their effect on the abundances is not mainly attributed to a galactic wind but to the gas infall. This could considerably amplify the tendency found in the presented cd results and lead to repeated element enhancement cycles in the N/O-O diagram. As a great success the cd models reproduce already convincingly the observed abundance peculiarities of dIrr in a self-consistent way. Since the cd treatment is appropriate for sensitively balanced systems of low gravitational potential energy like the disks of gSs and dIrrs, its models can provide a fundamental insight into interaction processes of small-scale energetics and large-scale dynamics. This has already led to a successful reproduction of metallicity distributions and the temporally coupled evolution of different regions in the Milky Way (Samland et al. 1997) by means of large-range streaming of hot gas and the mixing with locally infalling gas clouds. This same interaction of processes seems to affect also the evolution of dIrrs substantially. Acknowledgments: We are grateful to Mike Edmunds for helpful discussions. This work was supported by the Deutsche Forschungsgemeinschaft (DFG) under grant no. He 1487/5-3 and /23-1 (A.R.). The numerical calculation have been performed at the computer center of the University of Kiel and the NIC in Jülich. ## References Binggeli, B. 1994, in “Panchromatic View of Galaxies – their Evolutionary Puzzle”, eds. G. Hensler, J.S. Gallagher, C. Theis, Editions Frontières, Gif-sur-Yvettes, p. 173 Burkert, A. 1995, ApJ, 447, L25 Burkert, A., Hensler, G. 1989, “Evolutionary Phenomena in Galaxies”, Tenerife, eds. J.E. Beckmann & B.E.J. Pagel, Cambridge University Press, p. 230 Colless, M., Ellis, R.S., Taylor, K., Hook, R.N. 1990, MNRAS, 244, 408 Dekel, A., Silk, J. 1986, ApJ, 303, 39 Duc, P.-A., Mirabel, I.F. 1998, A&A, 333, 813 Ferguson, H.C., Binggeli, B. 1994, A&A Reviews, 6, 67 Franco, J., Cox, D.P. 1983, ApJ, 273, 243 Gallagher, J.S., Wyse, R.F.G. 1995, PASP, 106, 1225 Garnett, D.R. 1990, ApJ, 360, 142 Garnett, D.R., Skillman, E.D., Dufour, R.J., et al. 1995, ApJ, 443, 64 Grebel, E. 1997, Rev. Modern Astron., 10, 29 Guzman, R., Jangren, A., Koo, D.C., et al. 1998, ApJ, 495, L13 Hamann, F., Ferland, G.J. 1992, ApJ, 391, L53 Han, M., Hoessel, J.G., Gallagher, J.S., et al. 1997, AJ, 113, 1001 Heckman, T.M., Dahlem, M., Lehnert, M.D., et al. 1995, ApJ, 448, 98 Hensler, G., Rieschick, A. 1998, PASP Conf.Ser., 147, 246 Hensler, G., Rieschick, A. 1999, Proc. XVIIIth Rencontre de Moriond ”Dwarf Galaxies and Cosmology”, eds. T.X. Thuan et al., Editions Frontieres, Gif-sur-Yvettes, in press Hensler, G., Theis, C., Burkert, A. 1993, in Proc. 3<sup>rd</sup> DAEC Meeting “The Feedback of Chemical Evolution on the Stellar Content in Galaxies”, eds. D. Alloin & G. Stasinska, p. 229 Hensler, G., Theis, C., Gallagher, J.S. 1999, in preparation Ho, L.C., Fillipenko, A.V. 1996, ApJ, 472, 600 Hodge, P.W. 1989, ARA&A, 27, 139 Hunter, D., Wilcots, E.M., van Woerden, H., et al. 1998, ApJ, 495, L47 Izotov, Y.I. 1999, Proc. XVIII. Rencontre de Moriond, Les Arcs, “Dwarf Galaxies and Cosmology”, eds. T.X. Thuan, et al., Editions Frontieres, Gif-sur-Yvettes, in press Kobulnicky H.A. 1998, Proc. Quebec Conference: “Abundance Profiles: Diagnostic Tools for Galaxy History”, eds. D. Friedli et al., San Francisco: ASP, p. 108 Kobulnicky, H.A., Skillman, E.D. 1996, ApJ, 471, 211 Kobulnicky, H.A., Skillman, E.D. 1997, ApJ, 489, 636 Kobulnicky, H.A., Skillman, E.D. 1998, ApJ, 497, 601 Kobulnicky, H.A., Dickey, J.M., Sargent, A.I., et al. 1995, AJ, 110, 116 Kobulnicky, H.A., Kennicutt, R.C., Pizagno, J.L. 1999, ApJ, in press Kobulnicky, H.A., Skillman, E.D., Roy, J.-R., et al. 1997, ApJ, 477, 679 Köhler, S., Reimers, D., Wamsteker, W. 1996, A&A, 312, 33 Köppen, J. 1998, Java Applet ”http://www.astrophysik.uni-kiel.de/pershome/supas086/ChemFlows.html” Köppen, J., Edmunds, M. 1998, MNRAS, in press Köppen, J., Theis, C., Hensler, G. 1995, A&A, 296, 99 Köppen, J., Theis, C., Hensler, G. 1998, A&A, 328, 121 Koo, D.C., Guzman, R., Faber, S.M., et al. 1995, ApJ, 440, L49 Kunth, D., Matteucci F., Marconi G. 1995, A&A, 297, 634 Kunth, D. & Sargent W.L. 1986, ApJ, 300, 496 Larson, R.B. 1974, MNRAS, 169, 229 Lilly, S.J., Cowie, L.L., Gardner, J.P. 1991, ApJ, 369, 79 Madau, P., Ferguson, H.C., Dickinson, M.E., et al. 1996, MNRAS, 283, 1388 Maeder, A. 1992, A&A, 264, 105 Maeder, A., Meynet, G. 1989, A&A, 210, 155 Marconi, G., Matteucci, F., Tosi, M. 1994, MNRAS, 270, 35 Marlowe, A.T., Heckman, T.M., Wyse, R.F.G., Schommer, R. 1995, ApJ, 438, 563 Matteucci, F. 1986, PASP, 98, 973 Matteucci, F., Tosi, M. 1985, MNRAS, 217, 391 Meurer, G.R., Freeman, K.C., Dopita, M.A., et al. 1992, AJ, 103, 60 Meurer, G.R., Staveley-Smith, L., Killeen, N.E.B. 1998, MNRAS, 300, 705 Mirabel, I.F., Dottori, H., Lutz, D. 1992, A&A, 256, L19 Molaro, P., D’Odorico, S., Fontana, A., et al. 1996, A&A, 308, 1 Molaro, P., et al. 1998, MNRAS, 293, L3 Pagel, B.E.P. 1985, ESO workshop on ”Production and Distribution of C,N,O Elements”, ed. I.J. Danziger, F. Matteucci, K. Kjär, p. 155 Pagel, B.E.P., Tautvaisiene, G. 1998, MNRAS, 299, 535 Pettini, M., Lipman, K., Hunstead, R.W. 1995, ApJ, 451, 100 Pilyugin, L.S. 1992, A&A, 260, 58 Renzini, A., Voli, M. 1981, A&A, 94, 175 Rieschick, A., Hensler, G. 1999, in preparation Samland, M., Hensler, G., Theis, C. 1997, ApJ, 476, 277 Schaerer, D., Meynet, G., Maeder, A., et al. 1993, A&AS, 98, 523 Schaller, G., Schaerer, D., Meynet, G., et al. 1992, A&AS, 96, 269 Searle, L., Sargent, W.L. 1972, ApJ, 173, 25 Skillman, E.D., Melnick, J., Terlevich, R., et al. 1988, A&A, 196, 31 Smecker-Hane, T.A., Stetson, P.B., Hesser, J.E., et al. 1994, AJ, 108, 507 Theis, C., Burkert, A., Hensler, G. 1992, A&A, 265, 465 Thuan, T.X., Izotov, Y.I. 1997, ApJ, 489, 623 Thuan, T.X., Izotov, Y.I., Lipovetsky, V.A. 1997, ApJ, 477, 661 Tomkin, J., Lambert, D.L. 1984, ApJ, 279, 220 van der Hoek, L.B., Groenewegen, M.A.T. 1997, A&AS, 123, 305 van Zee, L., Salzer, J.J., Haynes M.P. 1998a, ApJ, 497, L1 van Zee, L., Westphal, D., Haynes M.P., Salzer, J.J. 1998b, AJ, 115, 1000 Vilchez, J.M. 1995 AJ, 110, 1090 Welch, G.A., Sage, L.J., Mitchell, G.F. 1998, ApJ, 499, 209 Williams, R.E., et al. 1996, AJ, 112, 1335 Wirth, A., Gallagher, J.S. 1984, ApJ, 282, 85 Woosley, S.E., Weaver, T.A. 1986, in “Radiation Hydrodynamics in Stars and Compact Objects”, eds. D. Mihalas & H.A. Winkler, Springer, Berlin, p. 91 Woosley, S.E., Weaver, T.A. 1995, ApJS, 101, 181
no-problem/9908/physics9908030.html
ar5iv
text
# Convergence of two-center expansions in positron-hydrogen collisions ## Abstract The positron-hydrogen atom scattering system is considered within the S-wave model. Convergence in the elastic scattering, excitation, ionization, and positronium formation channels is studied as a function of the number and type of states used to expand the total wave function. It is found that all unphysical resonances disappear only if near-complete pseudostate expansions are applied to both the atomic and positronium centers. preprint: submitted to Physical Review Letters; LB6633 In recent years there has been substantial progress in the field of electron-atom collisions. Close-coupling based methods, utilizing near-complete pseudostate expansions about the atomic center, have been shown to yield very accurate results for discrete excitation and ionization channels, see for example . This progress is primarily based on the study of the electron-hydrogen atom scattering S-wave model (only states with zero orbital angular momentum are retained) performed by Bray and Stelbovics . They showed, by simply taking a pseudostate expansion whose completeness improves with increasing number of states $`N`$, that cross sections for discrete and ionization channels converged at all energies. Pseudoresonances, typically associated with small $`N`$ calculations, disappeared for sufficiently large $`N`$, and convergence was to the correct independently evaluated results. This was an extraordinarily powerful result that is at the base of the substantial success recently enjoyed by the various implementations of the close-coupling theories. The situation for positron-atom scattering is somewhat more complicated. In addition to the atomically centered states the positronium formation channel needs to be included in the calculations in order to allow for all possible scattering channels. In other words, one needs a combined basis consisting of two independent basis sets. However, in this case if each of the components of the basis is large enough, one may expect instabilities in the calculations. The reason for this is that at small distances between colliding fragments functions of different basis components may essentially repeat each other due to their nonorthogonality. Thus, the close–coupling problem with the combined–basis–expansions is ill–conditioned. Is this an insurmountable obstacle? This question has remained unanswered for a long time. Mitroy, Berge, and Stelbovics and Mitroy and Ratnavelu have performed convergence studies for the full positron-hydrogen problem at low energies. Below the first hydrogen excitation threshold they showed good agreement between large pseudostate close coupling calculations and the highly accurate variational calculations of Humberston . However, at higher energies, particularly above the ionization threshold the situation is less clear. We adopt the often-used CC($`N,N^{}`$) notation for close-coupling calculations that utilise $`N`$ atomic eigenstates and $`N^{}`$ positronium eigenstates to expand the total scattering wave function. In addition, a bar, when applied to $`N`$ or $`N^{}`$ indicates that pseudostates rather than eigenstates are used. In 1991 Higgins and Burke showed that in the close-coupling calculation CC(1,1) a giant resonance appeared around 40 eV incident energy. Since that time a huge body of literature has been devoted to the study of this and other resonances above the ionization threshold, see the excellent review of Walters et al. . The positions and widths of the resonances have been studied extensively , even though, as noted by Walters et al. , the mid-seventies work of Simon shows that there may not be any positron-hydrogen scattering resonances above the ionization threshold. As far as we are aware the first convincing numerical evidence that shows disappearance of the above-threshold Higgins-Burke type resonances was given by Kernoghan, McAlinden, and Walters . They considered the full positron-hydrogen scattering problem using pseudostate expansions on both centers. Their 18-state CC($`\overline{9},\overline{9}`$) calculation included s, p, and d states for both centers. Thus it was not clear what was the primary reason, if any, for the disappearance of the Higgins-Burke type resonances. Moreover, at energies above the ionization threshold the authors encountered a new false pseudoresonance structure associated with the unphysical pseudostates. Smooth results for the total and the dominant partial $`1s`$ cross sections were obtained after applying an energy averaging procedure in order to remove the pseudoresonances. The unsmoothed partial $`2s`$ and $`2p`$ cross sections both for atom excitation and positronium formation contained significantly more pseudostructure . At the same time the convergent close-coupling theory (CCC) , which made use of CC($`\overline{N},0`$) calculations, i.e. neglected positronium formation, gave very good results for the total, elastic, excitation and ionization cross sections in energy regimes where positronium formation cross section is either zero or small. The CCC calculations showed no pseudoresonances. Based on these results Kernoghan et al (see also ) suggested using pseudostates only for hydrogen with a few eigenstates of positronium i.e., CC($`\overline{N},N^{}`$) type calculations. Results of a CC($`\overline{30},3`$) calculation with a 30–state hydrogenic pseudobasis of Bray and Stelbovics , supplemented by the three lowest positronium eigenstates showed a considerable improvement over the CC($`\overline{9},\overline{9}`$) results . Though the new basis did not completely remove false pseudostructure from the $`2s`$ and $`2p`$ positronium formation cross sections in the upper neighbourhood of the ionization threshold, the conclusion was that in two–centre scattering problems CC($`\overline{N},N^{}`$) type calculations are adequate. Since the works of Kernoghan et al and Mitroy pseudostate calculations of positron scattering off atoms have been performed mostly using bases of this type. The same conclusion became dominant in ion-atom collisions as well mainly due to the extensive investigations by Kuang and Lin . The purpose of the present Letter is to demonstrate a case where stable and convergent two–centre pseudosate calculations free of any pseudoresonances could be achieved only if near-complete pseudostate expansions are applied to both the atomic and positronium centers, i.e. where calculations of type CC($`\overline{N},\overline{N^{}}`$) are used. This is done for the simple S-wave model that retains only states of zero orbital angular momentum. The importance of the electron-atom scattering S-wave model, which continues to attract considerable attention , suggests that this model might also be useful in the positron-hydrogen case. However, for this case it has attracted little attention, with the CC(2,2) calculation by Mitroy being the biggest. Yet, this problem captures most of the difficulties associated with the full positron-atom problem just as the electron-atom S-wave model contains most of the difficulties associated with the full electron-atom scattering problems. At the same time it does not include the unnecessary, in the present context, generalities. Particularly, there is no a priori mechanism for any resonances including the Feshbach ones below the ionization threshold, due to the absence of states with non-zero angular momentum. This is why the model is ideal for convergence studies, as only smooth cross sections are expected. The convergent close-coupling method is extended by including the positronium formation channels. This extension is mainly based on the work of Mitroy . However, positronium formation matrix elements have been written as a coupling of 12 $`j`$-symbols resulting in only two–dimensional integrals and finite angular momentum sums. In addition, momentum–space pseudostates and corresponding formfactors are used in a compact analytical form. To evaluate the integral over the momentum of the virtual electron involving the logarithmic singularity special–purpose orthonormal polynomials have been calculated which yield an optimal Gaussian quadrature. Together, these features allow big pseudostate calculations to be performed efficiently. The generalized CCC computer code has been tested against the CC(3,3) calculations of Mitroy and Stelbovics and other momentum space close–coupling calculations. To evaluate the positron-hydrogen S-wave model the total scattering wave function is expanded in terms of bases consisting of two independent truncated Laguerre bases with corresponding exponential fall-off factors $`\lambda `$ and $`\lambda ^{}`$, leading to close-coupling calculations denoted by CC($`\overline{N},\overline{N^{}}`$). The $`\overline{N}`$ hydrogen (H) states were obtained by diagonalising the Hamiltonian with $`\lambda 1.0`$. The minor variation in $`\lambda `$ was made to ensure that the total energy $`E`$ was exactly half-way between two adjacent pseudothresholds, as the underlying integration rule requires . The positronium (Ps) states were obtained by diagonalising the corresponding Hamiltonian with $`\lambda ^{}=0.5`$. This choice for $`\lambda `$ and $`\lambda ^{}`$ results in approximately equal number of negative- and positive-energy states. No variation of $`\lambda ^{}`$ was performed due to the fact that at the larger energies (above 40 eV) considered here too large a variation would be necessary to ensure that $`E`$ is half-way between two adjacent Ps pseudothresholds. Instead, the CC($`\overline{N},\overline{N^{}}`$) calculations are performed for all possible $`E`$ that are half-way between two of the Ps pseudothresholds. Though in this case observables are calculated at predefined incident energies, to show convergence in the expected to be smooth cross sections, we combine the results for varying $`\overline{N}`$ and $`\overline{N^{}}`$. If convergence is obtained at the calculated energies such combined results should form smooth curves. To assure the convergence on both basis components we took $`\overline{N}=\overline{N^{}}`$. In general, the combined basis does not need to be symmetric in the number of the hydrogen and positronium pseudostates. The basic momentum space integral equations for the transition matrix elements have been solved using a 96-point Gauss quadrature at each incident energy. The accuracy of the solution of the integral equations has been carefully checked for the case of the largest basis. Any further increase in the number of quadrature points did not significantly change the results. The number of states $`\overline{N}=\overline{N^{}}`$ has been systematically increased from 1 to 17. In Fig. 1 the total cross section for the model is presented evaluated using a number of different bases. The CC($`\overline{N},N^{}`$) calculations were performed on a fine energy mesh allowing for representation as a continuous curve. The dotted curve labeled CCC is obtained from all of the CC($`\overline{N},\overline{N^{}}`$) calculations (one for each dot) with $`\overline{N}=\overline{N^{}}=11,\mathrm{},17`$, see above. For example, the results at around 114 and 39 eV are from the CC($`\overline{17},\overline{17}`$) calculation and the ones at around 101 and 36 eV are from the CC($`\overline{16},\overline{16}`$). Starting with the CC($`\overline{11},0`$) calculation we see a smooth cross section that is very large at small energies. Adding three Ps eigenstates ($`N^{}=3`$) results in a massive drop in the cross section at low energies (in fact $`N^{}=1`$ is sufficient for this drop), but leads to Higgins-Burke type resonances at around 15 and 50 eV. Adding a further two Ps eigenstates ($`N^{}=5`$) and increasing the atomic pseudostate expansion to $`\overline{N}=17`$ results in only a slight shift of the resonance to higher energies. The shift is due to the increasing number of Ps eigenstates as we found invariance in the CC($`\overline{N},3`$) cross sections for $`\overline{N}=11,\mathrm{},17`$. However, the CCC curve is smooth devoid of any resonance structure. Thus, within this model the only practical way to yield pseudoresonance-free cross sections is to utilize near-complete pseudostate expansions for both the atomic and Ps centers. We note that the elimination of the Higgins–Burke type resonances is achieved for relatively small $`\overline{N}=\overline{N^{}}=5`$, whereas pseudoresonance structure disappears from $`\overline{N}=\overline{N^{}}=11`$. Fig. 2 shows some of the individual components of the CCC total cross section given in Fig. 1. The most dominant is the elastic scattering cross section followed by the ionization and Ps formation cross sections. The Ps formation cross section is evaluated by summing the cross sections for the negative-energy Ps states projected onto the Ps true discrete spectrum. The ionization (breakup) cross section is evaluated by subtracting from the total cross section the Ps formation cross section and the sum of the atomic negative-energy state cross sections projected onto the true discrete atomic spectrum. Though the Ps formation cross section is generally very small the inclusion of the Ps channels considerably reduces the total cross section at energies below 70 eV (see Fig. 1). The excitation cross sections to $`n3`$ are given in Fig. 3. Above the Ps(1s) threshold of 6.8 eV the Ps(1s) cross section is dominant until around 20 eV where H(2s) becomes the largest inelastic cross section. Most importantly we see smooth results for all these cross sections even though their magnitude is quite small. This gives us confidence in the accuracy of the pesented results generally. In summary, the CCC method for electron-atom scattering has been extended to positron scattering with the inclusion of the Ps formation channel. This channel is particularly important for the S-wave model studied at the low and intermediate energies. Within this model the disappearance of the Higgins–Burke type resonances is only possible using pseudostate expansions on both the atomic and Ps centers. Other pseudoresonances, resulting from finite pseudostate expansions, disappear if sufficiently large expansions are taken. Utility of using simultaneously two near-complete expansions to yield convergent results at all energies of interest has been demonstrated. Taking as many as 17 of both atomic- and Ps-centered states represents some of the biggest calculations of this type ever performed, and suggests the utility of the present numerical implementation for full positron-atom scattering problems. The presented cross sections represent benchmark results that may be used for comparison with other theories. We suspect that CC($`\overline{N},\overline{N^{}}`$) type calculations will yield faster convergence and greater accuracy for the full positron-atom scattering system over the commonly-used CC($`\overline{N},N^{}`$) type calculations, particularly for the Ps formation cross sections. ###### Acknowledgements. Support of the Australian Research Council and the Flinders University of South Australia is acknowledged. We are also indebted to the South Australian Centre for High Performance Computing and Communications.
no-problem/9908/cond-mat9908166.html
ar5iv
text
# Aspects of the dynamics of colloidal suspensions: Further results of the mode–coupling theory of structural relaxation ## I Introduction The dynamics of suspensions of colloidal particles has been the topic of active research for many years . Whereas the motion of isolated Brownian particles has been well understood for long, less is known about the dynamics of concentrated suspensions. Direct particle interactions and solvent mediated hydrodynamic interactions (HI) are important if the colloidal volume packing fraction increases above a few percent . Experimental studies mainly employing dynamic light scattering (DLS) have provided a wealth of information on dense systems and are also the stimulus for the theoretical work presented in this contribution. The pioneering study by van Megen and coworkers of the liquid to glass transition in hard–sphere like colloidal dispersions has provided detailed data on the density fluctuations at this dynamic, ergodic–to–nonergodic transition . Besides their intrinsic interest, these experiments also made possible quantitative tests of predictions from the idealized mode–coupling theory (MCT) . Agreement of experiment and theory within errorbars of ca. 15 % has been reported. This comparison, which up to now has tested leading order asymptotic predictions and has thus restricted the validity of the theoretical results to small separations from the critical density, provides support for the glass transition scenario as described by MCT, which has also been studied for colloidal micronetwork spheres , charged colloidal systems , and colloidal emulsions . Recent DLS experiments by Segrè, Pusey and coworkers study hard sphere like systems at lower colloidal densities and report unexpected and seemingly unrelated scaling properties of the dynamic scattering functions . Thus the question arises for which density range below the glass transition the MCT describes the dominant physical mechanism observed in the dynamics of concentrated colloidal fluids and whether the reported scalings can be explained by MCT. Studies of a generalized Stokes–Einstein relation and optical measurements by Mason and Weitz further raise the question about the connection of the viscoelastic moduli to the collective and self particle motion at rather high densities which can also be considered using MCT . The MCT was developed starting from theories of the dynamics of simple liquids upon the realization that in that subsystem of the equations of motion which aimed at describing the structural relaxation there exists a bifurcation separating ergodic from non–ergodic motion . The physical mechanisms held responsible have been called “cage effect” and “back flow” phenomenon . This transition was suggested as origin of the slowing down and of the anomalies of the dynamics at the glass transition. The idealized MCT studies the structural relaxation neglecting all other, possibly present, long–time dynamical effects . The extended MCT discusses long–time ergodicity restoring corrections . The bifurcation at critical values of the thermodynamic parameters like the colloidal packing fraction, $`\varphi `$, introduces a small (separation) parameter, $`\epsilon =(\varphi \varphi _c)/\varphi _c`$, and the possibility of asymptotic expansions in $`\epsilon `$; see Refs. for references and detailed results. Two asymptotic scaling law regions can be shown. In the first, for intermediate times, the feedback mechanism of caging of particles, causes an ergodic–to–nonergodic bifurcation, which is characterized by universal power law decays. During the second, for longer times, the collective rearrangements of the cages requires cooperative dynamics, like the build up of back flow patterns first discussed for liquid helium . The strongly correlated dynamics manifests itself in a coupling of the time scales for this final process of the structural relaxation. As it describes, in the liquid, the decay of the incipient frozen glassy structure, it is not surprising, that its MCT description requires detailed information about the equilibrium structure. As the leading–in–$`\epsilon `$ asymptotic results exhibit numerous nontrivial universal features, experimental tests of MCT mainly address these and thus a number of corrections need to be considered: $`(i)`$ The structural relaxation itself leads to corrections of higher order in $`\epsilon `$ which limit the range of validity of the leading asymptotics. $`(ii)`$ The short–time or microscopic dynamics affects the transient to the structural dynamics and needs to be considered if no clear separation of time scales is possible. $`(iii)`$ Long–time relaxational mechanisms may be present, which bypass the structural relaxation, and lead to faster decay. The third correction appears to be absent in colloidal suspensions at the densities of interest for the present study and thus shall be neglected in the following . It is interesting to mention, though, that in colloidal emulsions droplet shape fluctuations cause long–time relaxation and can be explained within the extended MCT . Theoretical understanding of the first correction effect in lowest relevant order in $`\epsilon `$ has been achieved recently and is the basis of the present considerations. Lacking a deeper understanding of the microscopic transport effects of colloidal suspensions (point $`(ii)`$ above) a qualitative study shall be undertaken incorporating short–time effects with the most simple approximations compatible with the MCT description of the structural relaxation. Thus the limit of the predominance of the structural relaxation is estimated from numerical solutions of the MCT equations using simple models for the microscopic transient dynamics of colloidal suspensions. Brownian short–time diffusion with and without hydrodynamic interactions is considered . The paper is arranged as follows. Section II summarizes the equations of motion of the idealized MCT. Section III presents and discusses our results, focusing first on the aspects purely structural–relaxational and then on the microscopic transient effects. The mentioned experimental findings are addressed in section III.C. Short conclusions end the paper. ## II Equations of motion The idealized MCT of the liquid to glass transition leads to a closed set of non–linear equations of motion for the density fluctuations . Other dynamical variables are connected to them via the Zwanzig–Mori formalism and mode–coupling approximations. The theory aims at a description of the structural relaxation as it emerges from the (microscopic) short–time dynamics and slows down owing to the increasing density and thus increasing importance of particle interactions. The central quantity, the (normalized) intermediate scattering function, $`\mathrm{\Phi }_q(t)=\frac{1}{N}\varrho _q^{}(t)\varrho _q(0)/S_q`$, measures the time and wave vector dependence of the collective density fluctuations, $`\varrho _q(t)`$, around the average homogeneous density which for hard–sphere particles of diameter $`d`$ is converted to the volume packing fraction. The normalization is provided by the static structure factor, $`S_q=|\varrho _q|^2/N`$ . The equations of motion of the idealized MCT from which asymptotic analysis extracts the physically relevant long–time dynamics can be summarized as : $`\mathrm{\Phi }_q(t)m_q(t)+{\displaystyle \frac{d}{dt}}{\displaystyle _0^t}𝑑t^{}m_q(tt^{})\mathrm{\Phi }_q(t^{})=0,`$ (1) $`m_q(t)={\displaystyle \underset{𝐤+𝐩=𝐪}{}}V(q;k,p)\mathrm{\Phi }_k(t)\mathrm{\Phi }_p(t),`$ (2) where the coupling constants or vertices, $`V`$, are uniquely specified by the static structure factor, $`S_q`$; see Refs. for explicit formulae. Equation (2) approximates the autocorrelation function of the fluctuating forces by considering a force to arise between two density fluctuations interacting via an effective potential. Then the correlation function of the four density fluctuations is approximated by the squared density correlators and the effective potential enters the vertices. In Eqs. (1,2) the structural dynamics results from the equilibrium state of the fluid as captured in $`S_q`$ which is the only input. Thus the long–time structural relaxation of a dense suspension of interacting Brownian particles is predicted to be identical to the one of an atomic system if the interaction potentials of both systems coincide. Here we will consider hard sphere potentials only. By ansatz, MCT neglects the possibility of an ordered, crystalline state and thus in experimental comparisons crystallization has to be prevented. Then the structure factor $`S_q`$ of a liquid of hard spheres is known to be a regular function smoothly varying with packing fraction and consequently the vertices in Eq. (2) are regular functions of the single (for a liquid of hard spheres) thermodynamic state parameter $`\varphi `$. Brownian particles diffuse on (appropriately chosen) short distances and thus for short times . Incorporating this into MCT leads to the following simple model of colloidal suspensions close to the glass transition : $$\mathrm{\Phi }_q(t)m_q(t)+\frac{d}{dt}_0^t𝑑t^{}m_q(tt^{})\mathrm{\Phi }_q(t^{})=\frac{1}{q^2D_q^{\mathrm{s}.}}\dot{\mathrm{\Phi }}_q(t).$$ (3) This equation replaces Eq. (1) and extends it to short times where the initial condition runs $`\mathrm{\Phi }_q(t)\dot{=}1q^2D_q^{\mathrm{s}.}t`$. Two approximations for the short–time diffusion coefficient, $`D_q^{\mathrm{s}.}`$, are widely used in theoretical work on colloidal dispersions and differ in the treatment of the solvent effects. In the most simple model of Brownian diffusion the interaction of the solvent with the colloidal particles is modeled with a single friction coefficient, $`\zeta _0`$ . This leads to $`D_q^{\mathrm{s}.(\mathrm{B})}=D_0/S_q`$, where $`D_0`$ is given by Einstein’s law, $`D_0=k_BT/\zeta _0`$, and the denominator arises from particle interactions as was first argued by de Gennes in a related context . This approximation is not satisfactory except for very low packing fractions, because the solvent also leads to long–ranged and quasi–instantaneous interactions of the colloidal particles, called hydrodynamic interactions (HI). Whereas the HI do not affect the equilibrium statistics and thus $`S_q`$ of the colloidal system, their effects on short time scales cannot be neglected in general and are captured in a wave vector dependent amplitude, $`H_q`$ : $`D_q^{\mathrm{s}.(\mathrm{HI})}=D_0H_q/S_q`$. Progress on a detailed theory for $`H_q`$ has proven very difficult but has culminated in accurate results for it up to intermediate packing fractions . At the considered high packing fractions however, $`D_q^{\mathrm{s}.(\mathrm{HI})}`$ can only be estimated from experiments or simulations at present . Note, that our approach to incorporate HI into the MCT equations of motion only via $`D_q^{\mathrm{s}.}`$ differs from the one developed by Nägele and others , which aims at describing the dynamics at lower packing fractions, and which would affect the structural long–time dynamics. Our approach also differs from the work of Cohen et al. who incorporate aspects of the cage effect into an effective short–time $`D_q^{\mathrm{s}.}`$ . The role of the HI here also differs from a recent theory of Tokuyama et al. who consider the HI in nonequilibrated colloidal suspensions. Equations (2) and (3) have been solved repeatedly without HI and with different approximations for the structure factor of the hard–sphere fluid ; for the details of the numerical calculations see the quoted references. Various aspects of the known solutions will be connected to recent experimental observations in this contribution and new solutions taking HI into account via $`D_q^{\mathrm{s}.(\mathrm{HI})}`$ will be presented, which are an extension of the calculations in Refs. . Figure 1 shows the short–time diffusion coefficients entering the numerical calculations discussed in the following. The short time diffusion coefficient without HI, $`D_q^{\mathrm{s}.(\mathrm{B})}`$, follows immediately from the hard sphere structure factor, where the Percus–Yevick approximation is used . The $`S_q`$ shown also enter the vertices in Eq. (2). The short–time diffusivity with HI, $`D_q^{\mathrm{s}.(\mathrm{HI})}`$, is chosen as shown in Fig. 1. It is aimed at a discussion of the dynamics with HI transient to the structural relaxation, and thus, for the high densities considered, a rough approximation modeled from the experiments in is used. Outside the window $`0.77q/q_p2.4`$, $`D_q^{\mathrm{s}.(\mathrm{HI})}`$ is assumed constant for simplicity. The values of $`D_q^{\mathrm{s}.(\mathrm{HI})}`$ for $`q0`$, $`q=q_p`$ and $`q\mathrm{}`$ are adjusted to 0.2:1.74:1.0 mimicking the measured ratios . Within the mentioned wave vector window, the experimental data are modeled by: $`D^0/D_q^{\mathrm{s}.(\mathrm{HI})}=x(q/q_p)^2/\mathrm{ln}f_{q/q_p}^c`$, where $`f_q^c`$ is the MCT critical non–ergodicity parameter, and $`x=0.29`$ leads to a continous matching. “De Gennes narrowing” is present in $`D_q^{\mathrm{s}.}`$ in both approximations, as its inverse varies in phase with the structure factor $`S_q`$ and an appreciable slowing down of the short–time dynamics for wave vectors $`q`$ around the principal peak at $`q=q_p`$ results. Some representative numerical results for the collective density correlators, $`\mathrm{\Phi }_q(t)`$, obtained as specified in , are exhibited in Fig. 2, where also the shown wave vectors are indicated. The correlators of Fig. 2(a) correspond to a density rather close to the critical liquid–to–glass bifurcation point, $`\varphi _c=0.516`$, of this model. The reduced distance equals $`\epsilon =(\varphi \varphi _c)/\varphi _c=10^{n/3}`$, with $`n=9`$, where, as in the following, in order to simplify comparison with Refs. , the packing fractions will be reported by stating the number $`n`$. Decreasing the packing fraction to $`\epsilon =0.1`$ corresponding to $`n=3`$, results in the intermediate scattering functions of Fig. 2(b). Results with and without HI as shown in Fig. 2 for various packing fractions will be discussed in the following. Only the dynamics in the colloidal liquid phase is shown, where the correlators decay to zero during the final relaxation process, because the mentioned experimental studies focus on this final decay; for MCT results on hard sphere glasses at higher packing fractions see Refs. . It is also interesting to note that short range attractions can increase the colloidal glass packing fraction appreciably . ## III Results and discussion ### A Leading asymptotic scaling laws In lowest order in the separation parameter $`\epsilon `$, the MCT predicts the existence of two divergent time scales with two different scaling laws describing the dynamics in expanding windows in time or frequency; see the Refs. for detailed derivations and reviews of these results. In the first or $`\beta `$–scaling law window, a factorization property allows to separate the sensitive and rather universal dependences of the dynamics on the separation parameter and on time from the system specific dependences like spatial variation. $$\mathrm{\Phi }_q(t)=f_q^c+h_qG(t,\epsilon )\text{for }|\mathrm{\Phi }_q(t)f_q^c|1.$$ (4) The $`\beta `$–correlator is given by a homogeneous function, $`G\sqrt{|\epsilon |}g_\pm ^\lambda (t/t_\epsilon )`$, specified by one system specific parameter $`\lambda `$, which can be calculated for simple liquids from $`S_q`$ and determines all exponents of MCT . The first divergent scaling time $`t_\epsilon =t_0|\epsilon |^{(1/2a)}`$, lies in the center of the window of validity of Eq. (4) and, below the critical density, can be taken from the root of $`G`$: $`\mathrm{\Phi }_q(t=t_\epsilon )=f_q^c`$. The one parameter $`t_0`$, the crossover or matching time, remains as only remnant of the short–time or transient motion and can only be obtained from matching the asymptotic results to the full dynamics including some short–time model. In Fig. 2(a) one notices that $`t_0`$ differs by a factor $`1.2`$ for the two models of the transient. A shift of the curves with HI relative to the ones without HI collapses both sets of curves for times $`t>\mathrm{\hspace{0.17em}0.1}d^2/D_0`$. The great simplification of the dynamics provided by the factorization in Eq. (4) may be interpreted as resulting from a localization transition close to which density fluctuations relax via local rearrangements and not via mass–transport over larger distances. If the spatial variation of $`f_q^c`$ and $`h_q`$ is studied in detail , the localization may be traced back to the “cage–effect”, that particles are surrounded by next–neighbor shells whose ability to cage the central particles depends on the fluctuations of the local structure and thus, in a cooperative manner, on the dynamics of the caged particles themselves. In the second scaling law region, another set of divergent time scales, called $`\alpha `$–relaxation times $`\tau _q`$, appears and $`\alpha `$–master curves describe the final relaxation of the density correlators from $`f_q^c`$ to zero during that time window . $$\mathrm{\Phi }_q(t)\stackrel{~}{\mathrm{\Phi }}_q(t/t_\epsilon ^{})\text{for }\epsilon 0,t/t_\epsilon ^{}=\text{fixed}.$$ (5) This superposition principle states that the final relaxation processes (asymptotically) depend on the distance to the critical point only via the relaxation times $`\tau _q`$, which moreover are coupled, $`\tau _q=\stackrel{~}{\tau }_qt_\epsilon ^{}`$, and diverge upon approaching $`\varphi _c`$: $`t_\epsilon ^{}=t_0|\epsilon |^\gamma `$ with $`\gamma >(1/2a)`$. The equations, which the $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})`$ obey, are obtained in a special limit from Eqs. (1,2) and are consequently independent of the microscopic short–time dynamics . The resulting two–step relaxation scenario of the idealized MCT has been fully worked out for some simple liquids; see Refs. for calculations of the exponents and master functions of the two scaling regions for a hard sphere liquid. The quantities of most immediate interest to experimental observations of the final or $`\alpha `$–relaxation process are the relaxation times . Figure 3 presents results for the asymptotic $`\tau _q`$ from the model specified above and compares them to previous calculations using a different approximation (Verlet–Weis form) for the static structure factors of a hard sphere liquid . Very small differences in the $`\tau _q`$ result from the two approximations to $`S_q`$. A discussion of short–time sum rules for colloidal suspensions as done in the spirit of de Gennes leads to the prediction of (short–time) relaxation times obeying $`\tau _q^{\mathrm{s}.}=(1/q^2D_q^{\mathrm{s}.})`$. Such a behavior for the Brownian model, scaled to match $`\tau _q`$ for $`q=q_p`$, is also indicated in Fig. 3. The MCT $`\alpha `$–relaxation times obtained from Eqs. (1,2), where the transient does not enter, and the results from the short–time sum rules qualitatively are similar for not–too–small wave vectors because both vary in phase with the structure factor. Their different physical origins, however, clearly show up for small wave vectors where the short–time relaxation times, $`\tau _q^{\mathrm{s}.}`$, become diffusive, whereas the MCT $`\alpha `$–relaxation times $`\tau _q`$ become wave vector independent as first anticipated in Mountain’s description of Brillouin scattering in supercooled atomic liquids . Although the collective density fluctuations of the colloidal Brownian particles are diffusive on short time scales due to random collisions with solvent molecules, during the structural relaxation only stress fields arising from colloid–colloid particle interactions survive out to long times. Thus large distance density fluctuations decay by local particle rearrangements. The strong slowing down of $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})`$ on length scales of the order of the average next–neighbor distance indicates local and cooperative particle rearrangements and is reminiscent of the back flow phenomenon familiar from simple liquids . The coupling of the wave vector modes in Eqs. (1,2) explains the qualitative trend that the correlators with larger $`\alpha `$–process amplitudes, $`f_q^c`$, relax slower, i.e. have a larger $`\tau _q`$ . Intriguingly, for a hard sphere liquid the wave vector dependence of the dimension-less time scale $`\tau _q^{(\mathrm{f})}=r_s^2/(d^2\mathrm{ln}f_q^c)`$, is rather close to the one of the actual $`\alpha `$–time, $`\tau _q`$, at least for intermediate wave vectors. Note that the comparison shown in Fig. 3 must be taken with a grain of salt, as the definition of $`\tau _q`$ is not unique because of the stretching, i.e. non–exponentiality, of the $`\alpha `$–process in MCT. Nevertheless, this semi–quantitative connection of the $`\alpha `$–process amplitude to its time scale, suggests a possible (partial) collapse of the $`\mathrm{\Phi }_q(t)`$ for different $`q`$ at the same packing fraction onto a common curve given by: $$\mathrm{\Phi }_q(t)=\mathrm{exp}\{\frac{\mathrm{\Delta }r^2(t)}{6d^2\tau _q^{(\mathrm{f})}}\},\text{where }\tau _q^{(\mathrm{f})}=\frac{r_s^2}{d^2\mathrm{ln}f_q^c}.$$ (6) Conceptually, $`\mathrm{\Delta }r^2(t)`$ should be connected to the mean-squared displacement of a colloid particle, to be denoted by $`\delta r^2(t)`$. From the definition of $`\tau _q^{(\mathrm{f})}`$, Eq. (6), and the factorization property, Eq. (4), immediately follows that very close to the critical packing fractions all rescaled curves intersect at the $`\beta `$–scaling time $`t_\epsilon `$. The connection of $`f_q^c`$ via $`\tau _q^{(\mathrm{f})}`$ to the $`\alpha `$–relaxation time, see Fig. 3, then shepherds the correlators to stay close during the final relaxation step, too. Figure 4 shows representative scaled correlators for two packing fractions, where the used wave vectors are marked in the inset. The correlators are drawn for $`\mathrm{\Phi }_q(t)0.05`$ in order to prevent overcrowding the figure. The $`q`$–dependent stretching of the correlators causes a noticeable spreading of the rescaled correlators for long times. Considering Fig. 3, one also does not expect a data collapse for wave vectors outside the shown $`q`$–range. Moreover, this scaling explicitly violates the short time behavior of the intermediate scattering functions which e. g. become diffusive for small wave vectors invalidating Eq. (6). This explains the spread of the curves in Fig. 4 at short times. In Fig. 4 also the mean–squared displacement from Ref. shifted as suggested by Eq. (6) is shown and lies within the clatter of the curves. This ansatz, together with the known Gaussian approximation to the self intermediate scattering function , gives a most simplistic description of the coherent and incoherent density correlators of the MCT. Nevertheless, the only point where Eq. (6) asymptotically rigorously collapses all correlators is at $`\mathrm{\Phi }_q(t_\epsilon )=f_q^c`$ because of the factorization property, Eq. (4). Already in a vicinity of this point a spread of the curves exists because of $`h_qh_{\mathrm{MSD}}f_q^c/(6d^2\tau _q^{(\mathrm{f})})`$, which would follow from Eq. (6) and the known $`\beta `$–expansion, $`\delta r^2(t)\dot{=}r_{\mathrm{sc}}^2h_{\mathrm{MSD}}G(t)`$ . ### B Corrections The discussion up to now has used the asymptotic formulae to lowest orders in the separation parameter and thus might restrict the discussed phenomena to close neighborhoods of the critical packing fraction $`\varphi _c`$. The leading corrections in $`\epsilon `$ to the asymptotic scaling laws of Sect. III A have recently been discussed in detail for the present model , and in some cases allow to extend the range of validity of the asymptotic expansions appreciably. The corrections to the $`\beta `$–scaling law, Eq. (4), for the dynamics close to $`f_q^c`$ are of the form: $$\mathrm{\Phi }_q(t)=f_q^c+h_q(G(t)+H(t)+K_qG^2(t)+\epsilon \stackrel{~}{K}_q),$$ (7) where the $`K_q`$ and $`\stackrel{~}{K}_q`$ are wave vector dependent constants which follow from asymptotic solutions to Eqs. (1,2). See for the definitions and for the correction function $`H(t)`$, which is of order $`𝒪(\epsilon )`$. The range of validity of the $`\beta `$–scaling law, Eq. (4), is thus found to be of order $`\sqrt{\epsilon }`$, and to differ for different wave vectors or observables. The $`\beta `$–region description of Eq. (7) extends the range of usefulness of the MCT asymptotic expansion around the critical non–ergodicity plateau appreciably as can be seen in Refs. , and provides detailed few parameter formulae for the density correlators which have already found use in the data analysis of computer simulation studies . For the curves without HI of Fig. 2, Eq. (7) describes the correlators in the $`\beta `$–window on a 10% error level starting from the time scale $`t>\mathrm{\hspace{0.17em}0.1}d^2/D_0`$ which was estimated in Sect. III A to be the range of domination of the structural relaxation. Thus Eq. (7) extends the asymptotic expansions of the structural relaxation almost to the microscopic dynamics in this model. The $`\alpha `$–process has been the focus of the recent DLS scattering studies and, for wave vectors around the peak, describes the main portion of the decay of $`\mathrm{\Phi }_q(t)`$. The range of validity of the $`\alpha `$–process–superposition principle, Eq. (5), is appreciably larger than the one of the $`\beta `$–scaling law: $$\mathrm{\Phi }_q(t)\stackrel{~}{\mathrm{\Phi }}_q(t/t_\epsilon ^{})+\epsilon \stackrel{~}{\mathrm{\Psi }}_q(t/t_\epsilon ^{})\text{for }\epsilon 0.$$ (8) The corrections in Eq. (8) are only of linear order in $`\epsilon `$. Although the complete form of $`\stackrel{~}{\mathrm{\Psi }}_q(\stackrel{~}{t})`$ is not known yet, its variation for short rescaled times can be deduced and can be argued to give the dominant correction to Eq. (5) for not too large $`\epsilon `$: $`\stackrel{~}{\mathrm{\Psi }}_q(\stackrel{~}{t}0)=h_q\stackrel{~}{B_1}\stackrel{~}{t}^b`$, where the coefficient $`\stackrel{~}{B_1}`$ is of order unity. As this term can grow without bounds, the dominant aspect of the leading corrections is to cause the correlators $`\mathrm{\Phi }_q(t)`$ to rise above the $`\alpha `$–master curves for times shorter than the $`\alpha `$–relaxation time. In this time window, the $`\alpha `$–master curves follow von Schweidler’s law, $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})f_q^c=h_q\stackrel{~}{B}\stackrel{~}{t}^b`$ (again with $`\stackrel{~}{B}=𝒪(1)`$) . As the $`q`$–dependence of the time scales of the $`\alpha `$–correlators $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})`$ can be estimated from $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})\dot{=}f_q^c(1(t/\tau _q^{(\mathrm{vS})})^b)`$, with $`\tau _q^{(\mathrm{vS})}=(\stackrel{~}{B}f_q^c/h_q)^{1/b}t_\epsilon ^{}`$ , the short time corrections can be rewritten in the time window $`t_\epsilon tt_\epsilon ^{}`$: $$\mathrm{\Phi }_q(t)\dot{=}f_q^c(\mathrm{\hspace{0.17em}1}(t/\tau _q^{(\mathrm{vS})})^b\epsilon (\frac{t_\epsilon ^{}}{\tau _q^{(\mathrm{vS})}})^{2b}(t/\tau _q^{(\mathrm{vS})})^b).$$ (9) Thus it is apparent that deviations from the asymptotic $`\alpha `$–process scaling law, Eq. (5), are stronger for correlators with a shorter $`\alpha `$–relaxation time or smaller $`\alpha `$–process amplitude; the second connection arising because of the relation between $`\tau _q`$ and $`f_q^c`$, see Fig. 3. If at larger separations from the critical density the $`\alpha `$–relaxation times are determined from the correlators $`\mathrm{\Phi }_q(t)`$, then the corrections to the $`\alpha `$–scaling law, Eq. (8), may differently affect $`\tau _q(\varphi )`$. This is caused by the inherent stretching in the $`\alpha `$–master curves and by the time variation of the corrections $`\stackrel{~}{\mathrm{\Psi }}_q(\stackrel{~}{t})`$. The dominant short time variation of $`\stackrel{~}{\mathrm{\Psi }}_q(\stackrel{~}{t})`$ leading to Eq. (9) will affect $`\tau _q(\varphi )`$ if a definition of the relaxation times is used, which stresses the initial decay during the $`\alpha `$–process. A possible definition of $`\tau _q(\varphi )`$ exhibiting this effect is given by: $`\mathrm{\Phi }_q(t=\tau _q)=\frac{1}{2}f_q^c`$. Some results are indicated in Fig. 2, where also the $`\alpha `$–process amplitudes, $`f_q^c`$, are shown in the inset. As Fig. 5 shows, this definition of $`\tau _q`$ asymptotically gives almost identical $`q`$–dependences as obtained from the $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})`$ . Because of the rather large range of validity of the $`\alpha `$–scaling law, Eq. (5), for the intermediate scattering functions as explained by Eq. (8), only very small deviations of $`\tau _q(\varphi )`$ from the asymptotic wave vector dependence are seen in $`\tau _q(\epsilon )^\gamma `$ for $`n>3`$. As expected from Eq. (9), for $`n=3`$, which lies close to the limit of applicability of $`\alpha `$–scaling, the relaxation times are relatively longer and the largest (relative) deviations appear for correlators with small $`f_q^c`$ or $`\stackrel{~}{\tau }_q`$. At this separation from the critical density, $`\varphi =0.9\varphi _c`$, already some differences for the two models of the short time diffusion, with and without HI, are noticeable in Fig. 5. As shown in the inset, the differences can almost completely be incorporated into a packing fraction dependent shift of the matching time $`t_0`$. If the time scales are normalized to unity for $`q=q_p`$, then collapse can be achieved of the $`\tau _q`$ at $`n=3`$ except for the smallest wave vectors. Note that some finer aspects of the figure depend on the special choice how to measure $`\tau _q`$. For example, the correlators without HI at $`n=3`$ and for $`q=q_p`$ and $`q=0.94q_p`$ (just below it) actually almost overlap and the apparent differences in $`\tau _q`$ arise solely from the $`f_q^c`$–values entering the used definition. For even larger separations from the critical density, $`n=2`$ and $`n=1`$ in Fig. 5, clear differences of the long–time scales with and without HI appear and can obviously not be explained by structural relaxation, Eqs. (1,2), alone. The diffusive particle motion on short time scales causes the correlators for small wave vectors to decay slower relative to the non–diffusive $`\alpha `$–process. Often a diffusive behavior is assumed also for the structural relaxation and the relaxation times are converted to diffusion coefficients via $`1/D_q=q^2\tau _q`$. Fig. 6 shows so calculated $`D_q`$ normalized at $`q=q_p`$ in order to eliminate the drift of $`q_p`$ with packing fraction; see Fig. 1. Almost no deviations from the asymptotic variation as follows from the $`\alpha `$–scaling law, Eq. (5), can be be seen for $`n>3`$. Note that the non–diffusive character of the structural relaxation is hidden in this representation. For larger separations and thus smaller packing fractions, a trend of the long–time diffusion coefficients with hydrodynamic interactions (HI) to approach the shape of the short time ones, $`D_q^{\mathrm{s}.(\mathrm{HI})}`$, can be recognized. Considering Figs. 5 and 6, one needs to keep in mind, however, that differing methods to determine the final relaxation times or the long–time diffusion coefficients would lead to somewhat different $`q`$–dependences because they would weigh the stretching of the $`\alpha `$–process $`\stackrel{~}{\mathrm{\Phi }}_q(\stackrel{~}{t})`$ and the leading corrections $`\stackrel{~}{\mathrm{\Psi }}_q(\stackrel{~}{t})`$ differently. The definitions chosen here allow to explain the wave vector and packing fraction dependences in $`\tau _q`$ and $`D_q`$ from known aspects, Eqs. (8,9), of the asymptotic expansions. ### C Visco–elastic properties The time or frequency dependent shear modulus, $`G_\eta `$, of colloidal suspensions can be defined as an autocorrelation function of elements of the stress tensor and splits into three contributions . Whereas the first arises from the direct potential interactions of the particles and is familiar from simple atomic liquids, the latter two contain effects of the HI and are peculiar for colloidal particles immersed in a solvent. Only for the first potential part there exist MCT expressions which are applicable close to the glass transition at $`\varphi _c`$ ; however see Ref. for lower densities. Similarly as for $`m_q(t)`$ from Eq. (2), $`G_\eta (t)`$ is given by a quadratic mode–coupling functional in the $`\mathrm{\Phi }_q(t)`$. Consistent with the neglect of the HI contributions to $`G_\eta (t)`$, solutions for the $`\mathrm{\Phi }_q(t)`$ are used which are calculated without HI, i. e. with the short–time diffusion coefficients $`D_q^{\mathrm{s}.(\mathrm{B})}`$. Figure 7 shows the frequency dependent storage and loss shear moduli for a number of densities . As we consider the part of $`G_\eta `$ arising from potential colloidal interactions only, and thus cannot address the importance of HI at higher frequencies, only results in the frequency window of structural relaxation are shown. For low frequencies, the viscosity $`\eta `$ can be obtained via $`G_\eta ^{\prime \prime }(\omega 0)\omega (\eta \eta _{\mathrm{}})`$, where $`\eta _{\mathrm{}}`$ is the high–frequency shear viscosity which is caused by instantaneous solvent interactions . We use the approximation $`\eta _{\mathrm{}}=k_BT/(3\pi dD_0)`$. A plateau region in $`G_\eta ^{}(\omega )`$ corresponds to the $`\beta `$–scaling window, Eq. (4), and indicates elastic behavior of the colloidal suspension on intermediate time scales. In the non–ergodic states above $`\varphi _c`$, the colloidal system would be characterized by a finite elastic shear modulus, $`G_\eta G_\eta ^c`$, where the value at the glass transition follows from the $`f_q^c`$. The appropriate Fourier transforms of the $`\beta `$–correlator describe the dynamics around this elastic plateau in $`G_\eta ^{}`$, and in the minimum region of $`G_\eta ^{\prime \prime }(\omega )`$ between the transient high frequency dynamics and the $`\alpha `$–relaxation peak, which sensitively shifts with separation from the critical density. For a single colloidal particle in a continuum fluid the Stokes–Einstein relation connects the particle diffusion coefficient and the solvent viscosity, $`\eta D^{\mathrm{Self}}=\frac{k_BT}{3\pi d}`$. The self diffusion coefficient and the mean–squared displacement at finite colloid densities can, within MCT, be obtained from the autocorrelation function of the fluctuating forces which the single particle experiences from the colloidal liquid : $$\delta r^2(t)+D^{\mathrm{s}.\mathrm{Self}}_0^t𝑑t^{}m^{\mathrm{Self}}(tt^{})\delta r^2(t^{})=6D^{\mathrm{s}.\mathrm{Self}}t,$$ (10) where $`D^{\mathrm{s}.\mathrm{Self}}`$ is the short–time diffusion coefficient of the single particle which, neglecting HI, is given by $`D^{\mathrm{s}.\mathrm{Self}}=D_0`$ . The long–time self diffusion coefficient, $`D^{\mathrm{Self}}`$, follows from Eq. (10) in the Markovian limit, $`D^{\mathrm{s}.\mathrm{Self}}/D^{\mathrm{Self}}=1+D^{\mathrm{s}.\mathrm{Self}}_0^{\mathrm{}}𝑑tm^{\mathrm{Self}}(t)`$ . The memory function $`m^{\mathrm{Self}}`$ in MCT is given by another mode–coupling functional. Thus a priori, within MCT one would expect connections or similarities of $`m^{\mathrm{Self}}(\omega )`$ and $`G_\eta (\omega )`$ only because of the scaling laws. In the $`\beta `$–scaling region, asymptotically both functions exhibit the same shape , $`G_\eta ^{\prime \prime }(\omega )/h_{G_\eta }\chi ^{\prime \prime }(\omega )`$ and $`m_{}^{\mathrm{Self}}{}_{}{}^{\prime \prime }(\omega )/h_{m^{\mathrm{Self}}}\chi ^{\prime \prime }(\omega )`$, where $`\chi ^{\prime \prime }(\omega )`$ follows from the $`\beta `$–correlator $`G(t)`$ in Eq. (4). It is included in Fig. 7. The $`\alpha `$–superposition principle, Eq. (5), states that the $`\alpha `$–relaxation peaks in both functions asymptotically approach a density independent shape and shift in parallel upon varying $`\epsilon `$. This $`\alpha `$–scale coupling also immediately predicts the product $`D^{\mathrm{Self}}\eta `$ to approach a constant asymptotically for $`\varphi \varphi _c`$ . Nevertheless, as for example the $`\alpha `$–peak positions need not coincide, the close agreement of $`G_\eta (t)`$ and $`m^{\mathrm{Self}}(t)`$ in Fig. 7 over a wide window in time or frequency and covering a substantial variation in packing fraction is somewhat surprising. Presumably it arises, because, during the cooperative structural motion (cage effect), the collective density correlators around the peak in $`S_q`$, i.e. on the length scale of the average particle distance, dominate the dynamics of small–$`q`$ MCT memory functions. ### D Comparison with experiments The results of the MCT calculations of the previous sections, which partially have been tested in DLS experiments aimed at the glass transition , can also be used to discuss the recent experiments at somewhat lower densities which were mentioned in the introduction. Various other aspects of the results and their possible experimental relevance have been presented in and will not be repeated here. As a first aspect, let us point out, that if the mean–squared displacement can be measured and thus the connected memory function, $`m^{\mathrm{Self}}`$, then the numerical results show that a close estimate of the potential part of the shear modulus $`G_\eta `$ can be obtained. Even beyond the connections predicted by the two asymptotic scaling laws, Eqs. (4,5), the numerical results exhibited in Fig. 7, show that both functions are closely related, presumably because both arise from the cooperative cage dynamics. This connection may be considered as a frequency dependent generalization of the Stokes–Einstein relation and was assumed and tested in the recent diffusive wave spectroscopy measurements of Mason and Weitz . In another study of the same authors , they also observed that the $`\beta `$–correlators from Eq. (4) provide a description of the (directly measured) shear moduli spectra in an intermediate frequency window consistent with the MCT description of the potential part of $`G_\eta `$. The $`\alpha `$–scale coupling predicts that the various relaxation times and transport coefficients of a colloidal suspension close to the critical packing fraction $`\varphi _c`$ shift in parallel. For example, the prediction $`\frac{2k_BT}{\pi d\eta D^{\mathrm{Self}}}=`$ const. for $`\varphi \varphi _c`$ follows from Eq. (5). Quantitatively, the ratio approaches 5.93 , see Fig. 8, a value very close to the classical Stokes–Einstein prediction. Note however, that the conditions required for the classical Stokes–Einstein relation to hold, clearly are violated at packing fractions around the glass transition. A small but noticeable packing fraction dependence in $`1/(D^{\mathrm{Self}}\eta )`$ arises because of trivial density prefactors connecting the exhibited moduli of Fig. 7 with the transport coefficients. The $`\alpha `$–process corrections Eq. (8) and their discussion in Eq. (9) suggest that the $`\alpha `$–scale coupling should hold well for $`\alpha `$–relaxation scales obtained at low frequencies. This is supported by the observation, that the Stokes–Einstein relation considered with $`lim_{\omega 0}\varphi _cG_\eta ^{\prime \prime }(\omega )/(\varphi \omega )`$ replacing $`\eta `$, and $`m_{}^{\mathrm{Self}}{}_{}{}^{\prime \prime }(\omega =0)`$ replacing $`1/D^{\mathrm{Self}}`$, considerably reduces its density dependence. Upon decreasing the packing fraction to $`\varphi =0.9\varphi _c`$ ($`n=3`$), where the $`\alpha `$–scaling law looses validity, this ratio increases by 20 % relative to the asymptotic value, whereas the actual Stokes–Einstein ration increases by 29 %. Even larger density dependences can be expected if the long–time diffusion coefficients are obtained in time or frequency windows where the dominant corrections to the $`\alpha `$–process, see Eq. (9), appreciably increase the relaxation times of the correlators with shorter asymptotic $`\alpha `$–relaxation times. This effect is apparent in Fig. 8, where the wave vectors away from the peak position in $`S_q`$ show an increase in the relaxation times relative to the asymptotic $`\alpha `$–scale prediction, which on the other hand holds rather well for $`q=q_p`$. At $`q=1.17q_p`$ where $`S_q(\varphi _c)=0.90`$ a 65 % increase is seen at $`n=3`$, whereas at $`q=q_p`$ the Stokes–Einstein ratio changes only by 11 %. The results concerning the generalized Stokes–Einstein relation, which are presented in Fig. 8, and their explanations using Eqs. (8,9) rest on the simplifications caused by the bifurcation singularity in the MCT equations and by the entailing asymptotic expansions. As discussed, the range of validity of the $`\alpha `$–scale coupling does not appreciably exceed $`\epsilon 0.1`$ for the studied models of hard sphere like colloidal suspensions, and thus microscopic corrections to the MCT description of the structural relaxation need to be incorporated in principle beyond this distance to $`\varphi _c`$. Thus it is consistent with MCT that for packing fractions well below $`\varphi _c`$ the coupling of the time scales may continue in one system (hard spheres) but not in another one (charged spheres) . Also the estimate for the range of packing fractions where $`\alpha `$–scaling should hold can be expected to be model dependent, as the $`\alpha `$–master functions, Eq. (5), and their corrections, Eq. (8), depend on the fluid structure. Approximate expressions might be useful to describe the qualitative trends in the intermediate scattering functions, $`\mathrm{\Phi }_q(t)`$, in a wider context, in the same way as the Gaussian approximation is useful for the self intermediate scattering functions, $`\mathrm{\Phi }_q^s(t)`$. The Gaussian approximation was compared to the MCT results in . Recently Segrè and Pusey were lead by their DLS scattering data to propose such a formula and they observed partial collapse of their rescaled data. In Eq. (6) the possibility to collapse the intermediate scattering functions onto a common curve is studied using the dimensionless time scale $`\tau _q^{(\mathrm{f})}`$. This is suggested by the finding that the $`\alpha `$–relaxation amplitudes and relaxation times asymptotically are connected via $`\tau _q^{(\mathrm{f})}`$; see Fig. 3. The qualitative connection was expected but the quantitative closeness surprises and may be peculiar to the hard sphere system. Satisfactory data collapse using Eq. (6), see Fig. 4, is possible with deviations at long times because of the non–universality of the $`\alpha `$–process, and at short times, because Eq. (6) violates the short–time diffusive motion of colloidal suspensions. This reiterates that within MCT there is no connection of the obtained long–time diffusion coefficients, which follow from Eqs. (1,2), to the short–time ones . The only effect of the latter could be a shift in the time scale $`t_0`$, which matches the structural relaxation to the microscopic transient. Similar shapes, however, of the long– and the short–time diffusion coefficients with HI were observed by Segrè and Pusey in the recent DLS experiments on colloid fluids below and close to the glass transition . This similarity of the short– and long–time diffusion coefficients suggests to collapse the intermediate scattering functions with the assumption : $`\mathrm{\Phi }_q(t)=\mathrm{exp}\{(q^2/6)(D_q^{\mathrm{s}.}/D^{\mathrm{s}.\mathrm{Self}})\delta r^2(t)\}`$, which becomes exact for short times. In their experiments, Segrè and Pusey observed data collapse for wave vectors starting from somewhat below $`q_p`$ to the position of the second maximum in $`S_q`$. Figure 9 shows the solutions of the MCT equations with $`D_q^{\mathrm{s}.(\mathrm{HI})}`$ appropriately rescaled. Reasonable collapse of the curves onto a common one, which also is well represented by the mean–squared displacement, is observed in a similar wave vector range as in the experiments. For short times all curves coincide rigorously. For small wave vectors the non–diffusive character of the $`\alpha `$–process however leads to strong deviations for longer times. This trend also is present in the experimental data. The non–diffusive structural relaxation disagrees with the assumed diffusive scaling of the density correlators and thus cannot be rationalized with considerations of the short–time expansions following de Gennes. Unavoidable polydispersity effects in the experimental data could lead to additional deviations from the proposed scaling for small $`q`$, but no qualitative differences for samples of different polydispersities were reported in . Polydispersity effects could be incorporated into the present MCT following the work for charged colloids in Ref. . The partial collapse of the data for intermediate and long times results from the connections of the $`\alpha `$–process amplitudes to the time scales discussed in context with the ansatz of Eq. (6). Note that this connection may not be quantitatively satisfied as well in other colloidal systems, like e. g. charged colloidal particles . Thus the approximate scaling may hold less well in other systems. Differently from Eq. (6), the scaling with the short–time diffusion coefficients does not rigorously collapse the data at a longer time. But the close similarity of $`\tau _q^{(\mathrm{f})}`$ and $`D_q^{\mathrm{s}.(\mathrm{HI})}`$ explains that collapse for short times and approximate collapse for longer times is achieved. ## IV Conclusions In the idealized MCT, the long–time dynamics of colloidal liquids is dominated by the structural relaxation. Asymptotic expansions close to the critical packing fraction capture the qualitative aspects of the structural relaxation. In this contribution it is shown that the theoretical results also rationalize some recent experimental findings for larger separations from the critical density. Corrections to the coupling of the $`\alpha `$–relaxation times are wave vector dependent as seen in tests of generalized Stokes–Einstein ratios . The latter can be generalized to finite frequencies as observed in light scattering experiments , if the potential contribution to the shear modulus is considered. The tight coupling of the collective density fluctuations as captured in the scaling observed by Segrè and Pusey on the one hand supports the existence of an $`\alpha `$–scaling law as predicted by MCT, and on the other hand requires the quantitative connection of the $`\alpha `$–process amplitudes and relaxation times obtained here for hard sphere like colloidal particles. The failure of the scaling for small wave vectors seen in the experiments is predicted by MCT and highlights that the structural relaxation cannot be understood from short–time expansions. Such an approach, often referred to as “de Gennes narrowing” concept, would suppose diffusive colloidal dynamics for small wave vectors in disagreement with MCT and experiment . As the structural relaxation is determined by the equilibrium structure factor only, hydrodynamic interactions affecting the short–time and transient dynamics can be incorporated into the MCT without changing the long-time predictions. ## ACKNOWLEDGMENTS Valuable discussions with Dr. A. Latz, Dr. P. N. Segrè and Prof. W. Götze are gratefully acknowledged. This work was supported by the Deutsche Forschungsgemeinschaft under Grant No. Fu 309/2.
no-problem/9908/cond-mat9908221.html
ar5iv
text
# Coexistence of solutions in dynamical mean-field theory of the Mott transition ## Abstract In this paper, I discuss the finite-temperature metal-insulator transition of the paramagnetic Hubbard model within dynamical mean-field theory. I show that coexisting solutions, the hallmark of such a transition, can be obtained in a consistent way both from Quantum Monte Carlo (QMC) simulations and from the Exact Diagonalization method. I pay special attention to discretization errors within QMC. These errors explain why it is difficult to obtain the solutions by QMC close to the boundaries of the coexistence region. PACS numbers: 71.27.+a The Mott-Hubbard metal-insulator transition is one of the fundamental problems in the field of correlated electronic systems. During the last few years, dynamical mean-field theory (DMFT) has emerged as an appropriate paradigm for this transition. Within DMFT, the paramagnetic half-filled Hubbard model on the Bethe lattice was found to undergo a first-order phase transition at finite temperature $`T=1/\beta `$. The hallmark of such a transition is the coexistence of two solutions for the same value of the electronic interaction $`U`$ with $`U_{c_1}(\beta )<U<U_{c_2}(\beta )`$. As is usual, I will label “insulating” the solution which can be continuously followed at low temperature from large values of the interaction down to $`U_{c_1}`$ and “metallic” the one which exists from $`U=0`$ up to $`U_{c_2}`$. Recent work has challenged the above scenario. Using the Quantum Monte Carlo (QMC) method, rather than the Exact Diagonalization (ED) approach , the authors of were unable to detect coexistence in the DMFT equations. This is surprising, as the two algorithms were compared extensively to each other (cf, e.g. ref. , section VI, Appendix C). The aim of the present paper is to discuss the conflicting numerical approaches (for related work, cf ). I show that coexisting metallic and insulating solutions can indeed be obtained using both methods in a transparent and completely consistent way. I pay special attention to the problem of discretization errors. It will become clear that the QMC method needs to use very fine discretizations in order to obtain the insulating solution close to $`U_{c_1}`$. Inside the coexistence region, the problem disappears. All the numerical work in this paper relies on programs which are publicly available . Given the large number of articles already published on the subject, I will not repeat the standard definitions for the half-filled Hubbard model on the Bethe lattice within DMFT. These can be found e.g. in ref. . The important parameter of the model is the bandwidth $`D`$ (cf. ); I quote interactions and temperatures in units of the bandwidth $`D/\sqrt{2}`$. The point of departure of the present paper is ref. , where coexisting solutions of the mean-field equations at inverse temperature $`\beta D/\sqrt{2}=100`$ were found (within ED) for $`U_{c_1}<U<U_{c_2}`$ with $`U_{c_1}(\beta D/\sqrt{2}=100)3.3D/\sqrt{2}`$ and $`U_{c_2}(\beta D/\sqrt{2}=100)=3.8D/\sqrt{2}`$. For this paper, I choose as a reference point the value $`U=3.55D/\sqrt{2}`$, $`\beta D/\sqrt{2}=100`$. At this point, both metallic and insulating solutions can be found easily by ED . These solutions are given by bath Green’s functions $`𝒢_0(\tau )`$ and impurity model Green’s functions $`G(\tau )`$ (cf for definitions). As we are well within the coexistence region, both solutions correspond to the bottoms of deep basins of attraction. I have, e.g., perturbed both the metallic bath Green’s function, $`𝒢_0^{met}(\tau )`$, as well as the insulating one, $`𝒢_0^{ins}(\tau )`$, and fallen back into the corresponding solutions after a few iterations. This situation changes as we approach the boundary of the coexistence region. For smaller $`U`$, for example, the insulating solution becomes less and less attractive, until the basin of attraction vanishes at $`U_{c_1}`$. This scenario simply corresponds to the familiar free-energy landscape at a first-order transition. Both the ED and the QMC discretize some component of the DMFT equations. In ED, the discretization parameter is the number $`n_s`$ of sites of the quantum impurity model, while in QMC the number $`L`$ of slices appears, with $`\mathrm{\Delta }\tau =\beta /L`$ (cf. ). I have computed the insulating bath Green’s function $`𝒢_0^{ins}(\tau )`$ at the reference point for $`n_s=5,6,7`$. Each of the solutions (for fixed $`n_s`$) is fully converged, in addition, the convergence with the number of sites $`n_s`$ is excellent. I stress in passing that the sites in the ED algorithm are chosen in an optimal way, very similar to what is done in Gaussian integration . Exponential convergence in $`n_s`$ has been reported . Given that $`𝒢_0^{ins}(\tau ,n_s=7)`$ differs by less than $`0.005`$ from $`𝒢_0(\tau ,n_s=6)`$, I am led to the proposition that both $`𝒢_0^{ins}(\tau ,n_s=7)`$ and $`𝒢_0^{met}(\tau ,n_s=7)`$ are essentially exact. This assertion would have to be refuted for at least one of the solutions (in fact, for the insulating one) if we were to agree with the authors of ref. . The following reasoning will lead us to a very interesting result: consider a single iteration of the self-consistency loop $$𝒢_0^{ins,met}G𝒢_0^{\text{new}}$$ (1) If, contrary to our assertion, $`𝒢_0^{ins}(\tau )`$ or $`𝒢_0^{met}(\tau )`$ were not self-consistent, we should be able to detect differences, say, between $`𝒢_0^{ins}`$ and $`𝒢_0^{\text{new}}`$ \- even by other methods than ED, for example by QMC simulation. In this sense, I have discretized both $`𝒢_0^{ins}`$ and $`𝒢_0^{met}`$ and used them as inputs for single self-consistency loops (as in eq. (1)) of the QMC algorithm for different values of $`\mathrm{\Delta }\tau `$. From the discretized Green’s functions $`G^{\mathrm{\Delta }\tau }(\tau )`$, we then compute $`𝒢_0^{\text{new}}(\tau )`$ by inverse Fourier transformation. Results for the metallic solution are shown in Fig. 1. We see that the finite-$`\mathrm{\Delta }\tau `$ effects are quite small. Clearly, $`𝒢_0^{\text{new}}(\tau ,\mathrm{\Delta }\tau )𝒢_0^{met}(\tau )`$ as $`\mathrm{\Delta }\tau 0`$. The metallic solution has not been contested. The situation becomes much more illuminating as we consider the insulating solution $`𝒢_0^{ins}(\tau )`$. In this case, I have computed $`G^{\mathrm{\Delta }\tau }(\tau )`$ for $`\mathrm{\Delta }\tau =1,0.5,0.25`$. The results for $`𝒢_0^{\text{new}}(\tau )`$, again obtained by inverse Fourier transformation, are shown in Fig. 2: it is evident that very large finite-$`\mathrm{\Delta }\tau `$ effects are present. Notice that all the curves $`𝒢_0^{\text{new}}(\tau ,\mathrm{\Delta }\tau )`$ expose a plateau $`𝒢_0constant`$ for $`\tau `$ away from $`0`$ or $`\beta `$. Such a plateau is characteristic of an insulating solution (cf. Fig. 1), and its value decreases with the gap of the single-particle density of states. We thus arrive at the crucial observation that the finite-$`\mathrm{\Delta }\tau `$ effects bias the self-consistency condition of the QMC algorithm towards the metallic solution. Nevertheless, we find back the ED solution as $`\mathrm{\Delta }\tau 0`$. In terms of the free-energy landscape mentioned earlier, this simply means that the picture at finite $`\mathrm{\Delta }\tau `$ is tilted. The QMC iteration is dragged away from the insulating solution $`𝒢_0^{ins}`$ into the metal. Besides this effect, there seems to be no difference between the iteration loop of ED and the one of QMC. Within a deep basin of attraction of the insulating solution, a small drag due to finite $`\mathrm{\Delta }\tau `$ effects should only lead to a shift of the solution, and its stability should be preserved. This is exactly what I have observed. I have run full iteration loops starting from $`𝒢_0^{ins}(\tau )`$ for $`\mathrm{\Delta }\tau D/\sqrt{2}=1`$ and $`\mathrm{\Delta }\tau D/\sqrt{2}=0.5`$. In the first case, the simulation moved away from the insulation solution: After about $`10`$ iterations, the metallic solution was approximately recovered. In contrast, for $`\mathrm{\Delta }\tau D/\sqrt{2}=0.5`$, the insulating solution is very clearly stable. A $`40`$-day simulation of this single problem on a work station has obtained a very well-converged self-consistent solution for $`𝒢_0(\tau ,\mathrm{\Delta }\tau )`$ within QMC. This solution makes no more reference to ED, but it of course resembles the curves shown in Fig. 2; the plateau value is $`𝒢_0(\tau D/\sqrt{2}=50)0.30`$, comparable to the one-shot solution at the same value of $`\mathrm{\Delta }\tau `$. In my opinion, the present discussion of the discretization errors and both the one-step iteration and the explicit self-consistent Monte Carlo solution leave little room but to accept the coexistence at finite temperature. It is evident that the Monte Carlo simulation at the insulating solution has important finite-$`\mathrm{\Delta }\tau `$ effects. These errors modify the qualitative aspects of the solution only close to the phase boundaries where the basins of attraction of the insulating solution are shallow, and small. Therefore, the discretization errors will be more pronounced close to $`U_{c_1}`$, and also at higher temperature. This is what happened in the simulation of ref. , and what led the authors to wrong conclusions. Acknowledgement: it is a pleasure to thank G. Kotliar and A. Georges for helpful discussions.
no-problem/9908/astro-ph9908026.html
ar5iv
text
# A Shotgun Model for Gamma Ray Bursts ## 1. Introduction Gamma ray bursts (GRBs) show variability over a large range in time scales — from millisecond spikes in BATSE light curves (Bhat et al. 1992) to months in the associated afterglows (Bloom et al. 1999, Frail et al. 1997). The discovery of afterglow redshifts that place GRBs at cosmological distances implies isotropic energies of $`E_{\mathrm{iso}}10^{53}\mathrm{ergs}`$ for the average burst in gamma rays alone, released over an observed time of order $`T_{\mathrm{obs}}10\mathrm{s}`$. Since GRB spectra are believed to be optically thin, the only viable explanation is a highly relativistic outflow that releases its energy at sufficient distances from the central engine for the optical depth and the compactness of the plasma to be small, with required bulk Lorentz factors of $`\mathrm{\Gamma }100`$. In early models (M$`\stackrel{´}{\mathrm{e}}`$sz$`\stackrel{´}{\mathrm{a}}`$ros & Rees 1992, Katz 1994a, Sari & Piran 1995) the gamma rays were produced by the shock such an outflow drives into the interstellar medium (thus called ‘external shock models’). While providing a way to produce the required energy in gamma rays over the observed GRB durations, these models have not been very successful at explaining the observed short term variability: in order to see strong variability, the surrounding medium must be very clumpy (e.g., Shaviv & Dar 1995, Fenimore et al. 1996). It has been argued that this process might lead to very low efficiencies at converting kinetic into thermal energy (Sari & Piran 1997). However, a quantitative investigation casts some doubt on this argument (Dermer & Mitman 1999). These difficulties of the external shock model prompted Sari & Piran (1997) to postulate that the gamma ray emission must instead be produced by the internal shock scenario (Narayan, Paczy$`\stackrel{´}{\mathrm{n}}`$ski, & Piran 1992, Rees & M$`\stackrel{´}{\mathrm{e}}`$sz$`\stackrel{´}{\mathrm{a}}`$ros 1994). In this picture, the duration of the burst $`T`$ is set by the time scale over which the central engine operates, while the substructure in the bursts on time scales $`\tau `$ is produced by inhomogeneities in the outflow. These inhomogeneities are assumed to travel at different bulk Lorentz factors $`\mathrm{\Gamma }`$. Upon running into each other, these shells of material get shocked and release some of their kinetic energy in the form of gamma rays. This picture has become the paradigm in GRB physics. However, recent estimates of the energy conversion efficiency $`\eta `$ indicate that at best a few percent of the bulk kinetic energy carried by the outflow can be converted into gamma rays in internal shocks, which leads to uncomfortably high requirements on GRB energies (Panaitescu, Spada, & M$`\stackrel{´}{\mathrm{e}}`$sz$`\stackrel{´}{\mathrm{a}}`$ros 1999, Kumar 1999; larger efficiencies can be achieved if you assume large dispersion in $`\mathrm{\Gamma }`$, Katz 1997). A non-spherical geometry can reduce the required energy, however, a very small opening angle of the outflow implies a high rate of unobserved GRBs, which is hard to reconcile with the number of observed supernovae (which are believed to produce GRB precursors - either compact objects or hypernovae). See Piran (1999) for a detailed review of the subject. Fenimore et al. (1996, 1999, see also Woods & Loeb 1995) recently suggested that an external shock scenario could give rise to the observed variability if the spherical symmetry of the outflow were broken, still in the context of what Sari & Piran (1997) call a ‘Type I’ model, i.e., the burst duration is set by the slowing-down time of the ejecta. However, the observed temporal constancy of the pulse width in individual spikes of GRB990123 seems to rule out such a model (Fenimore, Ramirez-Ruiz, & Wu 1999. This would also argue against the scenario suggested by Dermer & Mitman, 1999.) We propose a different way by which substructure in the outflow can produce a GRB, also via the interaction with the external medium, but in a ‘Type II’ scenario, i.e., the duration of the burst is set by the lifetime of the central engine (see also Chiang & Dermer 1999, Blackman, Yi, & Field 1996). In this model, the outflow itself is very clumpy, with most of the energy concentrated in small blobs, which are sprayed out with high $`\mathrm{\Gamma }`$ over a small opening angle. These bullets then slam into the surrounding medium (not unlike a meteor shower or a shotgun blast), where they release their kinetic energy and produce gamma ray emission via external shocks, as described in §2. Note that, in this paper, we assume the existence of bullets; their possible origin is a topic for further research (see §3 for comments). Section 3 contains a discussion on radiative efficiencies, simulated burst light curves, and afterglows, and §4 contains a brief summary. ## 2. The Gamma Ray Burst Model BATSE light curves of most GRBs exhibit very spiky substructure (Fig. 1). In the context of our model, these spikes are produced by individual bullets of cold ejecta slamming into the surrounding medium. As we will show, a distribution of masses and/or Lorentz factors of these bullets can reproduce the observed signatures of GRBs reasonably well. In the following we assume that the central engine of the burst releases a number of bullets $`N^{}`$ distributed over a fan of opening angle of $`\theta 10^{}`$ with Lorentz factor $`\mathrm{\Gamma }1000\mathrm{\Gamma }_3`$ and released over a time period of $`T10100\mathrm{s}`$. Each bullet is assumed to freely expand sideways, with a sideways velocity of $`v_{}=\alpha /\mathrm{\Gamma }10^2\alpha _2/\mathrm{\Gamma }1/\mathrm{\Gamma }`$, measured in the observer’s frame ($`\alpha `$ is the sideways velocity in the comoving frame). The assumption that $`\alpha 1`$ implies that the internal expansion speed is very sub-relativistic. Since $`\theta 1/\mathrm{\Gamma }`$, we only see a fraction of the total released energy, $`E_{\mathrm{obs}}=10^{47}E_{53}/\mathrm{\Gamma }_{3}^{}{}_{}{}^{2}\mathrm{ergs}`$ (where $`E_{53}`$ is the inferred isotropic energy in units of $`10^{53}\mathrm{ergs}`$) and an observed number of bullets $`N100N_{100}=N^{}/(\theta ^2\mathrm{\Gamma }^2)`$. It is essential in our model that the covering fraction be less than unity (otherwise it would turn into an internal shock model), thus we require $`\alpha <1/\sqrt{N}`$. A bullet of mass $`M_\mathrm{b}=5.5\times 10^{13}E_{53}M_{}/N_{100}\mathrm{\Gamma }_{3}^{}{}_{}{}^{3}`$ will have converted half of its kinetic energy into internal energy (which can subsequently be radiated away as gamma rays, see below) after it has swept up or ploughed through a column of interstellar gas of mass $`M_\mathrm{s}M_\mathrm{b}/\mathrm{\Gamma }`$. If the duration of the observed spike is $`\tau _\mathrm{s}10^3\tau _3\mathrm{s}`$, then the length over which the material is swept up must be of the order of $`\mathrm{\Delta }R=3\times 10^{13}\tau _3\mathrm{\Gamma }_{3}^{}{}_{}{}^{2}\mathrm{cm}`$. If the ambient density is $`n_{\mathrm{amb}}10^8n_8\mathrm{cm}^3`$, the required Lorentz factor is $$\mathrm{\Gamma }=2300\left(\frac{E_{53}}{\alpha _{2}^{}{}_{}{}^{2}N_{100}\tau _{3}^{}{}_{}{}^{3}n_8}\right)^{1/8}$$ (1) This model fails for low ambient densities, as has already been discussed in the literature (e.g., Sari & Piran 1997). However, if the surrounding medium is very dense, $`n10^8\mathrm{cm}^3`$, Lorentz factors of $`\mathrm{\Gamma }1000`$ can explain the observed short term variability. The immediate conclusion is that in this model GRBs are not caused by mergers of naked compact objects. Rather, the required high ambient densities tie this model to the hypernova picture (Paczy$`\stackrel{´}{\mathrm{n}}`$ski 1998, Woosley 1993), which predicts that the material surrounding the blast is dense because of the pre-hypernova stellar wind. (Dense circum-GRB matter was also suggested in a different context by Katz, 1994b.) The fate of the outer layers of a hypernova precursor is unknown. If the bullets have to travel through a significant fraction of the star’s mantle (which is optically thick and thus not the site where the gamma rays are produced), their opening angle must be extremely narrow: $`\alpha <5\times 10^4E_{53}/(\sqrt{N_{100}}\mathrm{\Gamma }_3M_\mathrm{M})`$, where $`M_\mathrm{M}`$ is the mass of the mantle in units of $`M_{}`$. Since hypernovae are believed to originate from rapidly rotating massive stars collapsing into a compact object, it is possible that the material along the rotation axis has collapsed before the GRB, in which case the bullets would travel freely until they hit the circumstellar material. Similarly, little is known about the conditions of the pre-hypernova circumstellar material other than that it must be dense. Massive stars are known to have strong winds with mass loss rates of $`\dot{M}10^610^4M_{}\mathrm{yr}^1`$ and wind velocities from $`v_\mathrm{w}20\mathrm{km}\mathrm{s}^1`$ (red supergiant) to $`v_\mathrm{w}1000\mathrm{km}\mathrm{s}^1`$ (blue supergiant). These winds must still be present after the star collapses. In the following, we will assume that the GRB is produced in this leftover wind. The density profile in the ambient matter, then, roughly goes as $`nr^2`$ outside some radius $`R_0`$. It is natural to assume that $`R_0`$ is of the order of the stellar radius, $`R_010^{12}\mathrm{cm}`$ for a blue supergiant and $`R_010^{14}\mathrm{cm}`$ for a red supergiant. As a conservative estimate we assume that the sphere inside $`R_0`$ is evacuated. If we parameterize the density as $$n_{\mathrm{amb}}(r>R_0)=\frac{1.5\times 10^{36}\mathrm{cm}^3}{r^2}\frac{\dot{M}_6}{v_{20}}$$ (2) where $`\dot{M}_6`$ is the mass loss rate in units of $`10^6M_{}\mathrm{yr}^1`$, $`r`$ is in cm, and $`v_{20}`$ is the wind velocity in units of $`20\mathrm{km}\mathrm{s}^1`$, the observed slowing-down time scale is given by $$\tau =0.05\mathrm{s}\frac{E_{53}v_{20}}{\alpha _{2}^{}{}_{}{}^{2}N_{100}\mathrm{\Gamma }_{3}^{}{}_{}{}^{4}\dot{M}_6},$$ (3) independent of $`R_0`$. Thus $`\tau `$ can be of the order of a few milliseconds for both red supergiant and blue supergiant winds if $`\mathrm{\Gamma }1000`$. However, the observed time scale could conceivably be longer than this. The angular smearing time scale $`\tau _{\mathrm{ang}}`$ is defined as the spread in light travel time to the observer across the emitting surface. For a bullet at a viewing angle of $`1/\mathrm{\Gamma }`$ this gives $`\tau _{\mathrm{ang}}\alpha (R_0+\mathrm{\Delta }R)/\mathrm{\Gamma }^2`$, which is longer than $`\tau `$ if $`R_0\mathrm{\Delta }R/\alpha =1.5\times 10^{17}\mathrm{cm}E_{53}v_{20}/(\mathrm{\Gamma }_{3}^{}{}_{}{}^{2}\alpha _{2}^{}{}_{}{}^{3}N_{100}\dot{M}_6)`$. This is only of concern for very dense red supergiant winds, and only if the region interior to $`R_0`$ is evacuated. In Fig. 2 we show various limits on $`n_{\mathrm{amb}}`$ and $`\mathrm{\Gamma }`$ for a fixed opening angle of $`\alpha _2=1`$: * Each bullet is expected to plough through undisturbed medium. Thus, the covering fraction $`N\alpha ^2`$ of all the bullets together must be smaller than 1. For a slowing down time of $`\tau _3=1`$, this gives the dashed line in the plot, to the left of which the covering fraction is larger than unity. * The material between the location where the bullets release their energy and the observer must be optically thin. For $`\tau _3=1`$ and for $`R_0=0`$ (the most conservative limits) this constraint produces the dash-dotted line to the left of which the optical depth is larger than unity. * The forward shock must be radiative (see §3). This constraint is shown as a light grey region inside of which the shock is not radiative. * The angular smearing time $`\tau _{\mathrm{ang}}`$ must be smaller than the observed stopping time $`\tau `$. This limit is shown as a dotted line for $`R_0=10^{14}\mathrm{cm}`$. To the right of this line, the smearing time is longer than the observed slowing down time. The hatched region in the plot shows how the allowed region of parameter space opens up if we relax the time scale requirements to $`\tau _3=10`$. It is worth noting that this model makes an exception to the rule that external shocks cannot produce ‘Type II’ behavior (Sari & Piran 1997). This is for two reasons: First, the opening angle of the ejecta is so small that the angular smearing time is short compared to $`\tau `$. Second, the ambient density is so high that the observed slowing-down time is $`\tau 10^3\mathrm{s}`$. As a result, the total duration of the burst is determined by the time the central engine operates, while the short term variability is determined by the mass of the bullets and the statistical properties of the outflow. This is an important difference from the internal shock model, where the variability timescale is set by the intrinsic time scale of the central engine (e.g., the orbital time in a merger scenario). ## 3. Discussion In order for the efficiency of the burst to be reasonable, most of the internal energy produced in the shock must be radiated away immediately (this requirement holds for all GRB models). Electron synchrotron radiation is the only mechanism remotely efficient enough to produce the gamma rays. While the efficiency also depends on the transfer of energy from protons to electrons, we assume here that this process is efficient. Since we know the observed peak frequency of the gamma rays (of order $`500\mathrm{keV}`$, Piran 1999), we can then estimate the radiative efficiency under the assumption that the gamma rays are produced by electron synchrotron radiation. It is usually assumed that the magnetic field in the shocked gas is in equipartition with the energy density in relativistic particles. We therefore parameterize the magnetic field strength as $`U_\mathrm{B}ϵ_\mathrm{B}U_{\mathrm{B},\mathrm{eq}}`$, where $`U_{\mathrm{Beq}}`$ is the equipartition magnetic field energy density. Since the shocked wind material is likely flowing around the bullets at close to the speed of light (like a cocoon surrounding a radio jet), for efficient cooling we require that the cooling time scale in the comoving frame be smaller than the light travel time across the surface of the bullet $`\alpha (R_0+\mathrm{\Delta }R)/\mathrm{\Gamma }c`$. If that were not the case, the material pushed aside by the bullet would cool adiabatically rather than radiatively. Independent of $`R_0`$, this translates to the condition $`\mathrm{\Gamma }1430[\dot{M}_6\alpha _{2}^{}{}_{}{}^{2}ϵ_\mathrm{B}/v_{20}]^{1/4}\sqrt{E_\nu /500\mathrm{keV}}`$ (to the left of the light grey area in Fig. 2), where $`E_\nu `$ is the observed peak energy of the gamma rays. This is not a strict condition, however, since we do not know what the efficiency of the GRB is. Since our model assumes a central engine at work (essentially a black box shooting out bullets at a rate $`R(t)`$), any distribution of spikes could be reproduced, since $`R(t)`$ is arbitrary. It is, however, surprisingly simple to reproduce the main features seen in different burst profiles by varying only a few parameters in our model. For simplicity, we assume that all the bullets have the same initial $`\mathrm{\Gamma }`$ and the same sideways expansion rate, i.e., constant $`\alpha `$. We are left with two parameters — the number of bullets $`N_{100}`$ and the average slowing-down time (eq. ) — and two unknown functions: the mass distribution of the bullets $`N(M_\mathrm{b})`$ and the rate at which they are released $`R(t)`$. We assume that $`N(M_\mathrm{b})M_{\mathrm{b}}^{}{}_{}{}^{1/3}`$, chosen to give the observed power spectrum of $`P(\tau )\tau ^{1/3}`$ (Beloborodov, Stern, & Svensson 1999). For $`R(t)`$ we assume (for lack of better knowledge) that the bullets are released randomly over a time interval of 15 sec. To produce synthetic GRB lightcurves, we calculated the time dependence of $`\mathrm{\Gamma }`$ and the associated dissipation rate. Assuming the bullets are radiating efficiently and correcting for Doppler boosting and frequency shifts, we then computed the composite lightcurve for each bullet. We have plotted two simulated light curves in Fig. 1 for $`\tau _3=10`$ and $`N=100`$ (panel c) and $`N=10^4`$ (panel d). It seems that simply by varying $`N_{100}`$ and $`\tau _3`$ we can produce a wide range in light curve shapes. More complex features (like the gaps seen in panel a) must be related to the activity of the central engine and cannot be reproduced by a random spike rate as assumed above. We have also plotted the light curve produced by the deceleration of a single bullet along the line of sight (Fig. 1c, insert). Note that this profile is very similar to a true FRED (fast rise, exponential decay) profile. The rise is instantaneous and the decay follows a steep power-law (to first order). While indicative, these calculations are still rather crude and simplistic. A more careful analysis of shotgun GRB lightcurves should be carried out in the near future. Afterglows are an important test for any GRB model. How can we understand an afterglow in the context of the shotgun model? It is not immediately obvious why our model should produce an afterglow at all. This is because the bullets are assumed to spread sideways. As mentioned above, the bullets will have lost half of their kinetic energy at $`\mathrm{\Delta }R`$, where they have swept up $`1/\mathrm{\Gamma }`$ of their own mass. If we simply followed the dynamics of an individual bullet further in time, it would lose the rest of its energy exponentially quickly (Rhoads 1997). This is because the sideways velocity in the lab frame goes as $`1/\mathrm{\Gamma }`$, so that when $`\mathrm{\Gamma }`$ starts decreasing, the sideways velocity increases, which in turn increases the cross-sectional area of the bullet. As a result, the bullet can sweep up more mass, which leads to a run-away process. This would imply that the ejecta would come to a complete stop not far away from $`\mathrm{\Delta }R`$ and the observed afterglow would last less than a day. However, there are many bullets traveling together. As they expand, they start increasing the covering factor of the blast. Once the collection of bullets reaches unit covering factor, they stop slowing down exponentially, since further sideways expansion does not lead to an increase in swept up mass. As a result, the bullets start traveling collectively, resembling a collimated blast wave with opening angle $`\theta `$ rather than a meteor shower. The only possible difference between our model and the standard afterglow models is that in our model, the external density follows a power-law behavior instead of being constant, which has been discussed by Dai & Lu (1998). Since the opening angle of the merged blast is much larger than the opening angle of the individual bullets, the sideways expansion does not affect the dynamics until much later, when the blast has spread by an angle of order $`\theta `$ (Rhoads 1997). This transition from constant opening angle to rapid sideways expansion has been used in other models to explain the temporal break seen in the afterglow lightcurve of GRB990510 (Stanek, et al. 1999, Harrison et al. 1999, Sari, Piran, & Halpern 1999) It should be noted that we have not attempted to present a specific model for the production of bullets in this paper. We are confident that these bullets can be produced in GRB outflows and will concentrate on their production in future work. ## 4. Conclusions We have shown that GRBs can be produced by a shower of cold, heavy bullets shot at bulk Lorentz factor $`\mathrm{\Gamma }1000`$ into a dense medium. The required densities are consistent with a stellar wind from either a blue or red supergiant. This ties our model directly to the hypernova scenario. The gamma rays are produced by the shocks these bullets drive into the ambient gas. The total duration of the burst is then determined by the time the central engine operates rather than the slowing-down time of the bullets, while the latter produces the short-term variability seen in many bursts. After the gamma ray phase (when the bullets have lost half of their kinetic energy to radiation) the blast waves of the individual bullets merge into a single collimated shock front, which produces a standard afterglow in a declining external density profile. We have made no attempt to explain how a central engine might produce such a shower of bullets. This will be the subject of future work. This research was supported in part by NSF grants AST95–29170 and AST98–76887. MCB also acknowledges support from a Guggenheim Fellowship. We thank Pawan Kumar, Chris Reynolds, Mike Nowak, Jim Chiang, Martin Rees, and Annalisa Celotti for helpful discussions.
no-problem/9908/nucl-th9908023.html
ar5iv
text
# A next-to-next-to-leading-order 𝑝⁢𝑝→𝑝⁢𝑝⁢𝜋⁰ transition operator in chiral perturbation theory ## 1 Introduction Neutral pion production in proton-proton collisions $`pppp\pi ^0`$ near threshold needs no special introduction in the present section of these proceedings. The fact that the two dominant terms in the $`\pi N`$ hamiltonian, the P-wave $`\pi `$ emission from a nucleon and the Weinberg-Tomozawa term are forbidden by the selection rules demands an exploration of the less-well-known subdominant terms in the hamiltonian. One of these terms goes by the name of the Galilean term for an S-wave pion emission operator, Fig. 1(a) and is common to all nonrelativistic theories, as is the isospin-symmetric part of the elastic $`\pi N`$ scattering amplitude, Fig. 1(b). Since the exchanged $`\pi ^0`$ in Fig. 1(b) is off-shell, this graph is sensitive to the off-shell extrapolation of the $`\pi N`$ amplitude, which is not unique, however. All terms subdominant to the ones above are model-dependent. Precise cross section data measured in Bloomington and Uppsala are roughly five times larger than the simplest widely agreed-on prediction of the two diagrams in Fig. 1. One systematic proposal for corrections to graphs in Fig. 1 is based on the underlying chiral symmetry of the strong interactions and goes by the name of heavy-baryon chiral perturbation theory \[HB$`\chi `$PT\]. HB$`\chi `$PT generates additional subdominant terms in the hamiltonian that are clearly ordered by means of so-called chiral counting. This work is subject to the following assumptions: 1. Heavy-baryon chiral perturbation theory This means that we include: (i) only pion and nucleon degrees of freedom, i.e., no mesons heavier than the pion and no nucleon resonances; (ii) static nucleons in loop integrals and non-relativistic external nucleons, i.e., one can use the theory only near the threshold; (iii) well defined chiral order counting; this theory ought to provide smaller and smaller corrections to the leading-order result; (iv) perturbation theory, i.e., a finite number of graphs and counterterms, hence no dynamically created resonances or bound states. 2. Plane-wave initial and final states This means that we do not include the distortions of the initial and final waves, but give a brief discussion of the problems involved and of our present efforts to calculate the distorted wave effects. ## 2 Results We present the results of a systematic analysis of next-to-next-to-leading-order amplitude for the $`pppp\pi ^0`$ production at the threshold in heavy-baryon chiral perturbation theory. We found 19 topologically distinct new types of one-loop diagrams and one $`\pi `$ rescattering correction, that can potentially contribute to this reaction. The isospin selection rules and the S-wave character of the outgoing pion reduce the number of graphs from 19 to 6 (shown in Fig. 2). Analytic expressions for these effective transition operators are given in Ref. . In Table I we present our results on the relative importance of the six new types of diagrams, as well as the pion rescattering correction (type VII) that is topologically equivalent to the graph in Fig 1(b), as compared with the one-pion exchange rescattering graph in the plane-wave approximation for the initial and final states. The vertex-correction types of graphs, V and VI, are found to give only small corrections to the lower order result of Fig. 1(b), in conformity with expectations from general tenets of chiral perturbation theory. By contrast, we find very large contributions from the two-pion exchange graphs I - IV. Two of these types of graphs, II and III, cancel to within 1 % at the threshold, though they have very different energy- and momentum transfer dependencies at other kinematics, a point of some importance in distorted wave calculations. These “genuine” one-loop graphs can perhaps be interpreted as a part of an “effective $`\sigma `$-meson exchange” that is a significant part of an alternative theory based on the suggestion that the subdominant terms are induced by heavier-meson exchanges . The $`\pi `$ rescattering correction, type VII, is also large. ## 3 Discussion Appearance of large individual contributions calls for an explanation. Enhancement of the $`\pi `$ rescattering diagram VII over the leading-order graph Fig.1.b is perhaps the easiest to understand: The latter is proportional to the energy transfer between the protons $`k_0`$ which equals $`m_\pi /2`$ at the threshold, whereas the former is proportional to the three-momentum transfer $`\sqrt{\stackrel{}{k}^2}`$, which is determined by the kinematics of this reaction as $`\sqrt{\stackrel{}{k}^2}\sqrt{m_\pi m_N}\sqrt{7}m_\pi `$. The one-loop two-pion-exchange diagrams types I - IV, on the other hand, all involve one-loop integrals with two pion (and one nucleon) propagators $`I_{(\mu ;\alpha ,\beta )}(\omega ,v,P)`$ $`=`$ $`{\displaystyle \frac{1}{i}}{\displaystyle \frac{d^4l}{(2\pi )^4}\frac{(l_\mu ;l_\alpha l_\beta )}{(vl\omega i\epsilon )(m_\pi ^2l^2i\epsilon )\left[m_\pi ^2(lP)^2i\epsilon \right]}}`$ (1) In these loop integrals $`P`$ typically corresponds to the four-momentum transfer $`k`$ between the protons, which brings a factor $`k^2=m_\pi m_N`$ into the numerator accompanied by $`f_\pi ^2`$ coming from the vertices, thus resulting in a large enhancement factor, $`𝒪(k^2f_\pi ^2=m_\pi m_Nf_\pi ^215)`$. This explains the large size of the two-pion exchange diagrams, although their precise size is, of course, specific to the diagram. The deeper reason for the enhancement of both the loop graphs I - IV, and the rescattering graph VII lies in the derivative nature of the pion-nucleon couplings in $`\chi `$PT: it is the derivative coupling that puts the loop momentum $`l_\mu `$ into the numerator of the integral (1) and, as shown above, and it is the derivative coupling that makes the graph VII grow with the increasing three-momentum exchange $`\stackrel{}{k}`$. ## 4 Status of the distorted wave calculation It is known that the rescattering diagram Fig. 1(b) is sufficiently large to describe the data by itself, but it also interferes with the leading Born diagram Fig. 1(a). The latter gives a non-zero contribution only when evaluated between distorted waves, as four-momentum conservation prevents spontaneous pion emission. Consequently initial and final wave distortions are crucial to any numerical predictions of this reaction’s cross section. Preliminary work on the distorted-wave amplitude has led us to the following conclusions: (1) the $`\pi ^0`$ production operator is not square integrable: there are contact terms proportional to Dirac delta functions in configuration space. Consequently, the Fourier transforms from configuration to momentum space and vice versa are ill-defined; (2) the momentum space distorted wave integrals of the $`\pi ^0`$ production operator receive significant contributions from momenta up to 2 GeV, far beyond the region of applicability of HB$`\chi `$PT; (3) combined use of a $`\pi ^0`$ production operator that requires both energy and momentum transfer and two-proton Schrödinger wave functions that allow only three-momentum transfer can lead to inconsistencies. We wish to thank Shung-ichi Ando for pointing out an error in our calculation of the rescattering diagram, type VII.
no-problem/9908/astro-ph9908306.html
ar5iv
text
# Analytical Properties of the R1/m Law ## 1. Introduction After its introduction as a generalization of the $`R^{1/4}`$ law (de Vaucouleurs 1948), the $`R^{1/m}`$ law (Sersic 1968) has been widely used in observational and theoretical investigations (see, e.g., Ciotti & Bertin 1999, and references therein, hereafter CB). According to such law, the surface brightness profile is given by $$I(R)=I_0e^{b\eta ^{1/m}},m>0,$$ (1) where $`\eta =R/R_\mathrm{e}`$, and $`b`$ is a dimensionless constant such that $`R_\mathrm{e}`$ is the effective (half–luminosity) radius. The projected luminosity inside $`R`$ is given by $$L(R)=2\pi _0^RI(R^{})R^{}𝑑R^{}=I_0R_\mathrm{e}^2\frac{2\pi m}{b^{2m}}\gamma (2m,b\eta ^{1/m}),$$ (2) where $`\gamma (\alpha ,x)=_0^xe^tt^{\alpha 1}𝑑t`$ is the (left) incomplete gamma function. The total luminosity is then given by $$L=I_0R_\mathrm{e}^2\frac{2\pi m}{b^{2m}}\mathrm{\Gamma }(2m),$$ (3) where $`\mathrm{\Gamma }(\alpha )=\gamma (\alpha ,\mathrm{})`$ is the complete gamma function. It follows that $`b(m)`$ is the solution of the following equation: $$\gamma (2m,b)=\frac{\mathrm{\Gamma }(2m)}{2}.$$ (4) ## 2. Asymptotic expansion Unfortunately, eq. (4) cannot be solved in explicit, closed form, and so it is usually solved numerically. This is inconvenient for a number of observational and theoretical applications. At least three interpolation formulae for $`b(m)`$ have been given in the literature, namely $`b1.9992m0.3271`$ by Capaccioli (1989), $`b2m0.324`$ by Ciotti (1991), and $`b2m1/3+0.009876/m`$ by Prugniel & Simien (1997). These expressions provide an accurate fit in the range $`0.5m10`$; curiously, their leading term is linear in $`m`$, with a slope very close to 2. In CB it is proved that this behavior results from a general property of the gamma function. In particular, it turns out that the first terms of the asymptotic expansion of $`b(m)`$ as implicitly given by eq. (4) are $$b(m)2m\frac{1}{3}+\frac{4}{405m}+\frac{46}{25515m^2}+\mathrm{O}(m^3).$$ (5) Equation (5) now clearly explains the value of the interpolation formulae found earlier. Note that $`4/405=0.009876\mathrm{}`$. Of course, the asymptotic analysis provided in CB would allow us to give explicitly any higher order term if so desired. In CB it is shown that this expansion, even when truncated to the first four terms as in eq. (5), performs much better than the interpolation formulae cited above, even for $`m`$ values as low as unity, with relative errors smaller than $`10^6`$. The use of this simple formula is thus recommended both in theoretical and observational investigations based on the Sersic law. As a simple application, in CB the leading term of the asymptotic expansion of the total luminosity, of the central potential, and of the surface brightness profile associated with the R<sup>1/m</sup> law are obtained in explicit, analytical form. In particular, the connection of eq. (1) with the simple power law $`R^2`$, often used in the past to fit the photometric profiles of elliptical galaxies, is brought out explicitly. ### Acknowledgments. This work was supported by MURST, contract CoFin98. ## References Capaccioli, M. 1989, in The world of galaxies, H.G. Corwin & L. Bottinelli, Springer Verlag: New York, 208 Ciotti, L., 1991, A&A, 249, 99 Ciotti, L., & Bertin, G. 1999, submitted (CB) de Vaucouleurs, G. 1948, Ann.d’Ap., 11, 247 Prugniel, P., & Simien, F. 1997, A&A, 321, 111 Sersic, J.L. 1968, Atlas de galaxias australes. Observatorio Astronomico, Cordoba
no-problem/9908/hep-ex9908016.html
ar5iv
text
# 1 Summary of preliminary results ## 1 Summary of preliminary results The analyses designed for the 189 $`\mathrm{GeV}`$ data, described in detail in , have been applied without any modifications to the data collected by ALEPH up to July 27, 1999. The data sample comprises 29.0 $`\mathrm{pb}^1`$ at $`\sqrt{s}`$=191.6 $`\mathrm{GeV}`$ and 69.5 $`\mathrm{pb}^1`$ at $`\sqrt{s}`$=195.6 $`\mathrm{GeV}`$. The expected numbers of background events, the expected signal for a Higgs boson with $`m_\mathrm{h}`$=100 $`\mathrm{GeV}/c^2`$ and the numbers of selected candidates are listed in Table 1 for the hZ analyses and in Table 2 for the hA analyses. The reconstructed mass of the candidates selected by the NN- and cut-based hZ searches are displayed in Fig. 1(a) and (b). No departure from the Standard Model expectations is observed. This leads to improved 95% confidence level limits when the latest data are combined with those collected by ALEPH at lower energies. The confidence levels for the data in the background only hypothesis or the background and signal hypothesis are shown in Fig. 2(a) and (b). The limit obtained for the Standard Model Higgs boson is $`m_\mathrm{h}`$$``$98.8 $`\mathrm{GeV}/c^2`$, with an expected limit of 99.9 $`\mathrm{GeV}/c^2`$. The limit is calculated as described in , with the proper treatment of the overlap between the hZ and hA four jet and $`\tau ^+\tau ^{}`$ selections. The stand-alone hZ observed limit from the cut-based set of analyses is 97.6 $`\mathrm{GeV}/c^2`$, with 98.7 $`\mathrm{GeV}/c^2`$ for the expected limit. For the CP-even boson h of the MSSM, the limit is 85.2 $`\mathrm{GeV}/c^2`$, for all values of $`\mathrm{tan}\beta `$ above 1, with 86.1 $`\mathrm{GeV}/c^2`$ expected. Fig. 3 shows the excluded domain in the $`m_\mathrm{h}`$$`\mathrm{tan}\beta `$plane.
no-problem/9908/astro-ph9908004.html
ar5iv
text
# A case devoid of bias: ORS voids vs. IRAS voids ## 1 Introduction The Optical Redshift Survey (ORS; Santiago et al. 1995) and the Infrared Astronomical Satellite (IRAS) $`1.2`$-Jy redshift survey \[Fisher et al. 1995\] comprise an interesting data set pair: both are nearly all-sky surveys (excluding the Galactic plane region: $`|b|<20\mathrm{°}`$ for the ORS, $`|b|<5\mathrm{°}`$ for the IRAS) and as such are the densest, wide angle, three-dimensional samples currently available for study (at least until the forthcoming release of the IRAS PSC$`z`$ Redshift Survey, complete to 0.6-Jy – see Saunders et al. 1998). While sampling approximately the same volume, the two surveys differ vis-à-vis the galaxy populations they probe: IRAS and optical catalogues are each compiled using different selection criteria. As a result, IRAS galaxies are biased relative to optically selected galaxies – namely, they are less clustered and underrepresented in cores of galaxy clusters \[Strauss et al. 1992\]. As such, the two samples were already compared in several studies, e.g., the study of galaxy clustering and morphological segregation in the ORS \[Hermit et al. 1996\] and the derivation of the ORS-predicted velocity field \[Baker et al. 1998\], compared with the IRAS $`1.2`$-Jy gravity field \[Fisher et al. 1995\]. These works focused on properties derived from the distribution of galaxies in the surveys. In this study, we attempt a different approach, now comparing the distributions of voids in the two surveys. Voids are the most prominent feature of the large-scale structure of the universe, indeed occupying more than a half of the volume. Thus they are natural candidates for any quantitative large-scale structure study. We have already derived a void catalogue for the IRAS $`1.2`$-Jy survey \[El-Ad, Piran & da Costa 1997\], and in this paper we rederive a similar catalogue and present a new void catalogue for the ORS, using a suitably modified version of the void finder algorithm \[El-Ad & Piran 1997\]. But in addition to this additional void catalogue, the most interesting aspect of this paper is perhaps the comparison between the two void populations: if the galaxies in one survey are biased relative to the other, how does this affect the distribution of the voids? Being almost empty, and using a code which does its best at trying not to depend on the details of the galaxy distribution, voids could prove to be a relatively bias-free statistical probe. In this paper we derive in a consistent manner and compare void catalogues of the ORS and the IRAS $`1.2`$-Jy surveys. The paper is structured as follows. In §2 we describe the redshift catalogues we use, and in §3 we briefly review the void finder code and detail the modifications incorporated in order to analyze a survey with a non-isotropic selection function as the ORS. We then introduce the void catalogues (§4) and note some of the familiar voids we identify (§5). In §6 we compare the two void catalogues. Finally, in §7 we summarize our main conclusions. ## 2 The samples From the original redshift catalogues, we construct two semi–volume-limited samples with the same geometry: a sphere extending out to $`80h^1\text{Mpc}`$ with the volume-limited region comprising the inner $`r_0=50h^1\text{Mpc}`$. The Galactic plane is cut out of our samples eliminating the $`|b|<20\mathrm{°}`$ region, as we are limited by the wider ZOA of the ORS; hence our samples extend over 66 per cent of the skies. The volume examined is $`1.41\times 10^6h^3\text{Mpc}^3`$. The ORS catalogue \[Santiago et al. 1995\] contains over 8000 galaxies with redshifts, drawn from three sources – the UGC, ESO and ESGC catalogues. We choose to work with the diameter-limited ORSd sub-sample, as its sky coverage is wider than that of the magnitude-limited ORSm sub-sample (ORSm does not include the ESGC strip). After applying our geometrical cuts and volume-limiting, we end up with 2028 galaxies (Table 1 provides a break down of the galaxy counts per sub-catalogue: column 3 details the original ORSd catalogues, and column 7 details our final sample). The catalogue contains seven $`z`$-collapsed clusters. Note that since extinction corrections are properly taken into account, the volume limited region is not a perfect sphere – at directions where extinction is not negligible, the volume limited region is shallower having a depth $`r_0/10^{0.2\gamma A_B}`$. Here $`A_B(l,b)`$ is the extinction in the given direction and passband. $`\gamma `$ is the extinction correction parameter, for the ORSd being $`\gamma _d`$, the fractional decrease in isophotal diameter with extinction. We used $`\gamma _d=0.6`$ throughout \[Santiago et al. 1996\]. The various ORSd selection functions were derived as outlined in Santiago et al. \[Santiago et al. 1996\]. We use a parameterized form for the selection functions \[Yahil et al. 1991\]: $$\varphi (r)=\left(\frac{r}{r_s}\right)^{2\alpha }\left(\frac{r_{}^2+r^2}{r_{}^2+r_s^2}\right)^\beta $$ (1) where $`\varphi (r<r_s)=1`$ and $`\alpha `$, $`\beta `$ and $`r_{}`$ are free parameters whose best-fit values for our specific samples \[Santiago 1998\] are given in Table 1. The IRAS catalogue contains 5321 galaxies complete to a flux limit of 1.2-Jy \[Fisher et al. 1995\]. The sample we used, selected as explained above, has 1362 galaxies. It is important to note that the two catalogues are not independent – about half of the IRAS galaxies also appear in ORSd, and we have carefully taken this into account in our statistical analysis (§6). All of the analysis is performed in $`z`$ space. ## 3 Modifications in the VOID FINDER algorithm The void finder code used here to derive the void catalogues has been described in detail elsewhere \[El-Ad & Piran 1997\]. Briefly, the code covers voids using overlapping spheres, iteratively working its way starting from voids containing the largest spheres. Subsequent iterations identify new voids containing smaller spheres and improve the coverage of previously identified voids. Since voids need not be completely empty, an initial phase (wall builder) is used in order to filter out isolated galaxies which are allowed to be in the voids. Corrections are applied in the code in order to handle the observational selection function $`\varphi (r)`$. In a magnitude-limited sample, as we probe deeper we observe a smaller fraction of the galaxy distribution, hence the significance of finding an empty sphere declines with distance. The code corrects for this observational effect by weighing galaxies (during the initial phase) and spheres (during the construction of the voids) according to their distance. The ORS is more complicated to analyze than the previous surveys we have worked with (SSRS2s and IRAS $`1.2`$-Jy), since the usage of three different catalogues in its making (and the required extinction corrections) result in a non-isotropic selection function. We modified the void finder code in order to take this into account by appropriately weighing the galaxies and the spheres used to compose the voids. Each galaxy is assigned a weight \[Santiago et al. 1996\]: $$W_i=\frac{1}{\overline{n}_j\varphi _j(𝒓_i)}\frac{\overline{n}_{\mathrm{𝐼𝑅𝐴𝑆}_j}}{\overline{n}_{\mathrm{𝐼𝑅𝐴𝑆}}}$$ (2) where $`\overline{n}_j`$ is the mean number density of ORS galaxies in sub-catalogue $`j`$ in which galaxy $`i`$ happens to be located; $`\overline{n}_{\mathrm{𝐼𝑅𝐴𝑆}_j}`$ is the mean number density of IRAS galaxies inside that sub-catalogue; and $`\overline{n}_{\mathrm{𝐼𝑅𝐴𝑆}}`$ is the total mean number density of IRAS galaxies (see Table 1, column 10). The selection function of each sub-catalogue $`\varphi _j(r_i)`$ is usually just a function of the distance $`r`$, but in order to take extinction into account we adjust it: $$\varphi _j(𝒓_i)=10^{0.2\gamma A_B}\varphi _j(r_i)$$ (3) where $`A_B(l,b)`$ are the direction-dependent absorption coefficients \[Burstein & Heiles 1982\]. Consequently, we calculate a weight for each sphere considered to be a part of a void by volume averaging over the weights of points within each sphere: $$\overline{W}(𝒓_{\mathrm{centre}},d)=\left(\frac{4\pi }{3}d^3\right)^1_{\mathrm{volume}}W(𝒓)$$ (4) for a sphere centred on $`𝒓_{\mathrm{centre}}`$ with radius $`d`$. In order to estimate the underdensity of the ORSd voids we derive $`n_{1j}(r_0)`$, the selection function based galaxy number density for each sub-catalogue $`j`$ up to $`r_0`$. The prescribed number density is $`n_{1j}(r_s)`$ where $`r_s=5h^1\text{Mpc}`$ (see Table 1, column 8), and we calculate $`n_{1j}(r_0)=n_{1j}(r_s)\varphi _j(r_0)`$ (column 9). The actual galaxy number density for $`r<r_0`$ in our samples agrees well with the $`n_{1j}(r_0)`$ values, except for ESOd where the actual number density is significantly higher due to the presence of four nearby clusters (Doradus, Hydra, Centaurus and Fornax). See Fig. 1. For the purpose of calculating the void underdensities: if a void extends over several sub-catalogues, we derive the underdensity in each part of the void separately, and then volume-average the partial underdensities; and if a void extends beyond $`r_0`$ we weigh the galaxies in it using the relevant catalogue’s selection function. Note that since the calculation is done separately in each sub-catalogue $`j`$ relative to $`n_{1j}`$, the weight now is simply $`1/\varphi _j`$ and no relative density corrections (as in Equ. 2) are required. So the underdensity of a void with volume $`V`$ containing galaxies at locations $`𝒓_i`$ is: $$\frac{\delta \rho }{\overline{\rho }}=\underset{j}{}f_j\left(\frac{_i\varphi _j^1(𝒓_i)}{V}/n_{1j}(50)1\right)$$ (5) where $`f_j`$ is the fraction of the void that happens to be in sub-catalogue $`j`$. To estimate the voids’ statistical significance we use our usual confidence level \[El-Ad & Piran 1997\]: $$p(d)=1\frac{N_{\mathrm{Poisson}}(d)}{N(d)}$$ (6) where $`N_{\mathrm{Poisson}}(d)`$ is the number of voids in a Poisson distribution that contain a sphere whose diameter is $`d`$, and $`N(d)`$ is the same quantity for an actual survey. Poisson distributions are constructed using the same luminosity functions and extinction coefficients as the survey they correspond to. Our quoted confidence level should not be confused with the usual $`\sigma `$ grade; as such, it is a rather conservative grade since it does not take into account the total volume of a void, but rather only the size of the largest sphere that fits into it. Our $`p`$ is based on this aspect of the voids since it is the size of the largest sphere within a void that triggers a void’s initial identification by the void finder. ## 4 The void catalogues ### 4.1 ORSd The wall builder identified 1909 (94 per cent) of the galaxies as wall galaxies which may not reside in voids. Of the remaining 119 galaxies, 100 were found to be in voids (see Table 3). We identified 19 voids in the ORSd for which $`p>0`$; of these 9 have $`p>0.95`$, and we list these in Table 2: Column (1) identifies the voids with index numbers. Column (2) indicates $`p`$, the confidence level. Column (3) lists the diameters of equal-volume spheres; the volumes are tabulated in column (4). Column (5) lists the distance to the void centres, and the centres locations are detailed, in supergalactic coordinates, in column (6)–(8). Column (9) lists the void underdensities. Column (10) indicates the matching void(s) in our IRAS void catalogue (see §4.2), and column (11) measures the fit between the corresponding voids (see Equ. 7 in §6). Finally, column (12) identifies some of the familiar voids we identify (see §5). The 9 ORSd voids with $`p>0.95`$ occupy 46 per cent of the survey’s volume; an additional 8 per cent are occupied by the remaining 10 voids, bringing the void filling factor to $`0.54`$. The average equivalent diameter of the 9 significant voids is $`\overline{d}_{\mathrm{ORSd}}=49h^1\text{Mpc}`$ and the average underdensity in these voids is $`\overline{\delta }_{\mathrm{ORSd}}=0.89`$. ### 4.2 IRAS In the IRAS, $`92.5`$ per cent of the galaxies were located in walls, and 5 per cent in the voids (see Table 3). The void finder identified 16 voids with $`p>0`$ in the IRAS; of these, the initial 11 correspond to the above mentioned 9 significant ORSd voids (see Table 2, column 10). These 11 voids occupy 43 per cent of the volume (with the additional 5 voids occupying 5 per cent), with $`\overline{d}_{\mathrm{𝐼𝑅𝐴𝑆}}=44h^1\text{Mpc}`$ and $`\overline{\delta }_{\mathrm{𝐼𝑅𝐴𝑆}}=0.88`$. ## 5 Cosmography Two of the significant voids (4 and 7; and perhaps also 5) and most of the $`p<0.95`$ voids identified here are new and are not listed in the literature. Familiar voids which were already listed elsewhere are indicated in column (12) of Table 2. Voids 1–3 are all located within the volume of space probed by the Southern Sky Redshift Survey (SSRS; da Costa et al. 1998). We have already compiled a void catalogue of the southern Galactic cap $`m_B15.5`$ edition of this survey (SSRS2s), and we identify void 1 with EPdC voids 6+7; void 2 corresponds to EPdC void 5 \[El-Ad & Piran 1997\]. Void 2 is also known as the Sculptor void, and was identified in an earlier SSRS paper \[da Costa et al. 1988\] as SV3. Void 3 was identified in that paper (see Table 1 there) as SV2. Void 9 is pointed out in the recently published south Galactic cap $`m_B15.5`$ CfA survey (CfA2s; Huchra, Vogeley & Geller 1999). Void 8 corresponds to the large void found in the $`26\stackrel{}{.}5<\delta <32\stackrel{}{.}5`$ slice of the (north) CfA2 survey \[de Lapparent, Geller & Huchra 1986\]; this void is also identified (marked V4) in Table 1 of Saunders et al. \[Saunders et al. 1991\]. The voids listed by Tully \[Tully 1986\] are mostly beyond the range of our sample, but it is likely that voids 5 and 9 are the nearby tips of Tully voids 1 and 2, respectively (see Table 1 in the above referenced paper). Tully’s Local Void \[Tully 1987\] is defined to cover the region closer than $`30h^1\text{Mpc}`$ in the approximate direction (in Galactic coordinates) $`|b|<60\mathrm{°}`$ and $`0\mathrm{°}<l<90\mathrm{°}`$. We find in this direction voids 4, 8 (north of the Galactic plane), 3 and 9 (south of the Galactic plane) – though they all lie deeper than $`30h^1\text{Mpc}`$. The cosmographical tour of Strauss & Willick \[Strauss & Willick 1995\] mentions several other voids identified here: the void indicated to lie between the Local and Coma superclusters is void 6. The void beyond the Virgo cluster is void 11 and the void in the foreground of the Perseus-Pisces supercluster is void 19. The latter two voids have $`p<0.95`$ and thus are not listed in Table 2, but they can be viewed in Fig. 2. ## 6 Discussion The two void images are similar, although the two galaxy samples are quite different. In Fig. 2 we present two sets of slices covering the supergalactic plane and slices immediately above and below it, all together spanning $`15<Z_{\mathrm{SG}}<15`$. Beyond the visual impression, we quantify the spatial similarity between the two void populations by deriving $`\eta `$, the ratio between the overlapping volume of the two void distributions and their union: $$\eta =\frac{V_1V_2}{V_1V_2}$$ (7) where $`V_1`$ represents the volumes occupied by voids in one distribution, and $`V_2`$ is the same quantity for the other distribution. In case of exact overlap we have $`\eta =1`$; for two random distributions the expected overlapping volume is simply $`f_1f_2`$, the product of the two void filling factors. E.g., if $`f_1=f_2=0.5`$ the expected score for a pair of random samples is $`\eta =1/3`$. On the other hand, if 80 per cent of the volumes overlap we would get $`\eta =2/3`$. See Fig. 3 for an illustration of the void overlap in the supergalactic plane. Serendipitously, the details for our case happen to be quite similar to the above example: $`f_{\mathrm{ORSd}}=0.54`$ and $`f_{\mathrm{𝐼𝑅𝐴𝑆}}=0.48`$. The overlap score is $`\eta _{\mathrm{ORSd}\mathrm{𝐼𝑅𝐴𝑆}}=0.58`$. In contrast, for random ORSd- and IRAS-like samples where half of the random IRAS locations are used in the random ORSd distributions we get $`\eta _{\mathrm{rnd0}.5}=0.19\pm 0.04`$; for completely unmatched sample pairs the score is $`\eta _{\mathrm{rnd}}=0.13\pm 0.02`$. The theoretical expectation for $`\eta _{\mathrm{rnd}}`$, based on the filling factors of the random samples, is $`0.10\pm 0.02`$, and the small actual excess over it is likely due to the geometrical constraints of the sample. The above test measures the overall correlation between two void distributions without trying to match individual voids. As we identified (by eye) the corresponding voids in the two surveys (Table 2, column 10), we can also measure how well do the individual voids overlap. We do this by deriving $`\eta `$ values for a void from one of the samples and its counterpart(s) in the other sample. We report $`\eta `$ values for the ORSd–IRAS pairs in Table 2, column (11). The average value is $`\overline{\eta }=0.47`$. We can also give this score a more intuitive interpretation, by converting it to $`d_{\mathrm{misfit}}(\eta )`$, the distance (in fractions of a diameter units) at which one would need to place two identical spheres in order to get a specific value of $`\eta `$. Hence if there is exact overlap $`\eta =1`$ and $`d=0`$. If there is no overlap $`\eta =0`$ and $`d>1`$. A good reference point is at $`d=0.5`$ (two identical spheres misplaced by one radius), for which $`\eta =5/27`$. The average result we got for the ORSd–IRAS pairs corresponds to $`d_{\mathrm{misfit}}(\overline{\eta })=0.25`$. Note that difference in void volumes is a contributing factor in the derivation of the (mis)fit scores. E.g., for two spheres at the same center but with one radius being $`0.57`$ of the other we would get the same $`\eta `$ as for identical spheres one radius apart. As on average the ORSd voids are somewhat bigger than the IRAS voids, we can quantify the contribution of this factor to the misfit score: it translates to a base misfit value of $`d_{\mathrm{misfit}(\mathrm{volume})}0.1`$. In addition to the spatial correlation, the two distributions are also similar with regard to the average void properties – the total void filling factor, the average void diameter $`\overline{d}_{\mathrm{void}}`$ and the average void underdensity $`\overline{\delta }`$ (see Table 3). We find the later similarity to be of special interest: estimates of the relative bias between optical and IRAS samples based on the distribution of galaxies find $`(b_{\mathrm{opt}}/b_{\mathrm{𝐼𝑅𝐴𝑆}})^{\mathrm{gal}}1.5`$ \[Lahav, Nemiroff & Piran 1990, Baker et al. 1998\]. However, as the galaxy underdensity in the voids in both surveys is $`\overline{\delta }0.9`$, the void finder analysis shows that on average there is practically no biasing in the voids: $`(b_{\mathrm{opt}}/b_{\mathrm{𝐼𝑅𝐴𝑆}})^{\mathrm{void}}=1`$. The only other work we are aware of which compared optical and IRAS galaxies in voids examined the Boötes void \[Dey, Strauss & Huchra 1990\]. There it was found that the density contrast of IRAS galaxies within the Boötes sphere is roughly equal to the (optical) upper limit for that region \[Kirshner et al. 1987\]. In this work we examine a distribution of voids, and use many more galaxies ($`100`$, compared to 12 IRAS galaxies in the Boötes); still, our result of little – or no – biasing between optical and IRAS galaxies in the voids is consistent with the Boötes result. ## 7 Summary In this paper we present a comparison between two void distributions. These distributions sample the same volume of space, but were derived using the void finder code from two different galaxy samples – chosen optically (ORSd) and by the IRAS. The 9 significant voids we find in the ORSd match very well the locations of their IRAS counterparts, and our overlap/union ($`\eta `$) test shows a correlation significantly in excess of random. Combined with our previous analysis of the SSRS2s sample, we now have 3 different void catalogues all showing similar void properties, including the filling factor, average equivalent diameter and underdensity. In all our samples so far voids are limited by the boundaries of the surveys – in this paper, by the ZOA and the limited depth ($`80h^1\text{Mpc}`$); and in the SSRS2s by the narrow declination span (37$`\stackrel{}{.}`$5). In order to overcome this limitation we intend to further extend our void catalogues using deeper (LCRS – Shectman et al. 1996) and wider (CfA2s – Huchra, Vogeley & Geller 1999) samples. The fact that we find practically no biasing in the voids $`(b_{\mathrm{opt}}/b_{\mathrm{𝐼𝑅𝐴𝑆}})^{\mathrm{void}}=1`$ indicates that voids may be a relatively bias-free environment. As such, they comprise an attractive target with which one can examine different cosmological models, and we intend to explore this possibility using $`N`$-body simulations. Acknowledgments. We are indebted to Basilio Santiago for his help in unraveling the mysteries of the ORS. We thank Ofer Lahav, Myron Lecar, David Meiri and Sune Hermit for helpful discussions and comments. HE was supported by a Smithsonian Predoctoral Fellowship.
no-problem/9908/astro-ph9908214.html
ar5iv
text
# The Evolution of 3CR Radio Galaxies from z=1 ## 1 Introduction It has been known since the early 1980’s that the hosts of powerful radio galaxies display a tight relation between the $`K`$-band magnitude and redshift (Lilly & Longair 1984). In recent years it has been shown that, at least out to $`z=1`$, essentially the same relation is followed by both the less powerful Parkes Selected Regions (PSR) and 6C radio galaxies (Dunlop et al. 1989; Eales et al. 1997) and by X-ray selected brightest cluster galaxies (Collins & Mann 1997). When the $`K`$-band magnitudes of the 3CR galaxies are corrected for the expected effects of passive evolution of their stellar populations they appear to represent rather good standard candles from the present day out to redshift one and greater. Consequently, for many years it was widely accepted that the tightness of the $`Kz`$ relation for the 3CR radio galaxies could be most naturally explained by the most powerful radio galaxies having a rather well-defined mass, being formed at $`z1`$, and evolving basically passively thereafter. However, this long-held view has recently been challenged in a series of papers by Best, Longair & Röttgering (1997,1998), hereafter BLR. Between 1994–1996 BLR undertook an extensive study of a virtually complete sample of $`28`$ powerful FRII 3CR galaxies in the redshift range $`0.6<z<1.8`$ (see Fig 1). BLR obtained V and I–band HST images, J and K–band UKIRT images, and complimentary radio observations at 8.4 GHz with the VLA. BLR made use of the four broad–band images they obtained for each of their sources to perform spectral synthesis fitting. The four broad–band fluxes where fitted by a simple two-component model consisting of an old stellar population and a power-law contribution which represented any possible aligned component. The stellar population SED’s were constructed from the models of Bruzual & Charlot (1993), and assumed a 1-Gyr burst of star formation at $`z_{for}=10`$, with the stars evolving passively from that point onwards to the redshift of the particular 3CR galaxy. The assumed form of any possible nuclear/aligned emission was a power-law of the form $`f_\nu \alpha \nu ^{0.4}`$ which was used as a compromise between the likely contributions of quasar/starburst/dust scattered emission. The results of this fitting procedure (BLR 1998) show that the broad–band fluxes from these 3CR galaxies are fully consistent with that expected from an old passively evolving stellar population, as required in the traditional interpretation of the $`Kz`$ relation. However, the results of the determination of the elliptical galaxy half-light radii $`(r_e)`$ performed by BLR, combined with a summary of the literature concerning the environments of 3CR galaxies at high and low redshift, led BLR to challenge the traditional view of the $`Kz`$ relation. Using the method of fitting de Vaucouleurs templates to azimuthally averaged luminosity profiles, BLR produced best–fit scalelengths for 19 of their 28 objects (excluding 3C41 and 3C22 which have a large nuclear contribution (Leyshon & Eales 1998) not dealt with in the BLR fitting scheme). The results of this profile fitting revealed a mean scalelength of $`r_e=14.7\pm 1.3`$ kpc ($`\mathrm{\Omega }_0=1,H_0=50`$). As pointed out by BLR this is significantly larger than the $`r_e=8.2\pm 1`$ kpc found to be typical of low redshift ellipticals by Schombert (1987), and only around a factor of two smaller than the average scalelength of low-$`z`$ brightest cluster galaxies (BCG), $`r_e=32.7\pm 1.1`$ kpc, found by the same author. Combining this result with the $`L\alpha r_e^{0.7}`$ relation found by Kormendy (1977), and confirmed by our HST AGN host galaxy study (Dunlop et al 2000), BLR argue that even at $`z=1`$ the hosts of 3CR radio galaxies are highly evolved massive systems, possibly larger by a factor of two than their low-$`z`$ counterparts. Combining the apparent similarity of the $`z=1`$ 3CR galaxies to BCG at the same redshift, with results from the literature concerning the environments of high redshift radio galaxies (see BLR 1998 for a review), BLR concluded that the passive appearance of the $`Kz`$ relation is a cosmic conspiracy. The general picture given by existing studies is that the environments of $`z1`$ 3CR’s are consistent with moderately rich clusters; galaxy–galaxy cross-correlation function measures are consistent with Abell class 0, or richer (Hill & Lilly 1991), while multi-colour imaging shows many companion objects consistent with passively evolving coeval galaxies (Dickinson 1997). As pointed out by BLR, the same is not found in existing studies of the environments of low redshift 3CR’s (Prestage and Peacock 1988), where it has been reported that these objects reside in fairly low-density environments and have smaller characteristic sizes and luminosities than their high-redshift counterparts (Lilly and Prestage 1987). On the basis of the apparent redshift dependence of 3CR cluster environments BLR argued that the 3CR galaxies at $`z=0`$ and $`z=1`$ cannot be linked by a simple “closed-box” passive evolution scheme. In the picture favoured by BLR, the $`Kz`$ relation can be explained by the two 3CR galaxy samples having different dynamical evolutionary histories, but being observed at the points in their respective histories where they both contain a characteristic mass of stars; a few times $`10^{11}\text{ }\mathrm{M}_{}`$. The $`z1`$ 3CR galaxies would reach this mass first due to their location in a larger density peak producing strong merging between $`z=31`$. The 3CR galaxies at low-$`z`$ would reach the characteristic mass later through merging within their weaker cluster environment between $`z=1`$ and the present day, as is expected in many hierarchical galaxy formation models (Kauffmann 1999). The 3CR galaxies at $`z=1`$ would obviously also grow through merger activity in this redshift range and would become the radio dormant BCG we see today at low redshift, radio activity having ceased presumably due to the lack of available gas to feed their central black-hole. Although this picture is both consistent and appealing, it is dependent on the acceptance of environmental studies which now date back $`1520`$ years, together with the finding that 3CR galaxies at $`z1`$ have sizes comparable to BCG at that epoch, something which is not found to be true for low-$`z`$ 3CR galaxies. It was a careful examination of the methods used by BLR in determining their scalelengths which provided the original motivation for this paper. As mentioned above, scalelengths were presented by BLR for 19 of their 28 object sample (the images of the excluded 9 objects having poor signal-to-noise or suffering from extreme aligned emission). However, only 6 of these 19 scalelengths where actually determined from the high resolution HST imaging. The reason for this is that the one-dimensional profile fitting technique employed was unable to cope with the masking of significant contamination by companion objects or aligned emission. The majority of the scalelength information presented by BLR was in fact derived from the analysis of their $`K`$-band imaging data obtained at UKIRT. As reported by BLR, this imaging was obtained in $`1^{\prime \prime }`$ seeing conditions with typically 54 minutes of on-source integration which, considering the redshift range of the objects, leaves a relatively small amount of data which has both good signal-to-noise and is free from substantial seeing effects. The use of an analytical gausian function to represent what has proven to be the extremely complicated IRCAM3 PSF (McLure et al 1999b) casts further doubt on the quoted scalelength values. Experience gained from the comparison of our modelling results of $`z0.2`$ AGN host galaxies (McLure et al 1999a, McLure et al 1999b) with those previously obtained from typically $`1^{\prime \prime }`$ $`K`$-band imaging (Dunlop et al. 1993, Taylor et al. 1996) has demonstrated to us that the twin benefits of high resolution and a temporally stable PSF tend to combine to revise the best-fitting galaxy scalelength downwards. The main reason for this systematic trend appears to be that the high spatial resolution provided by HST allows the reliable identification of separate companion objects and their subsequent masking from the modelling process, preventing the inclusion of extra flux from biasing the scalelength determination to higher values. This effect is obviously only going to be strengthened at redshifts of $`z1`$ where the angular separation between companion objects and the target can easily be of the same order as the typical seeing experienced during the BLR UKIRT observations. Given these considerations it was felt worthwhile to re-examine the publically available HST data to determine whether the use of the full two-dimensional analysis utilised in the $`z0.2`$ quasar host galaxy programme would produce significantly different scalelength figures from those published by BLR. ## 2 Data Reduction All the HST images taken of the 3CR galaxies in the 28-object BLR sample were obtained from the HST archive facility<sup>1</sup><sup>1</sup>1http://archive.stsci.edu. A detailed list of redshifts, filters and exposure times can be found in Table 1 of BLR (1997). An investigation of the shorter-wavelength exposures of each object (mainly F555W, F622W) confirmed that they were either of insufficient signal-to-noise to be useful, or dominated by emission aligned with the radio axis. Of the 28 objects in the sample, preliminary analysis of the $`I`$-band images revealed a total of 16 which could be successfully modelled with our two-dimensional technique. The observational parameters of these 16 objects are detailed in Table 1 and illustrated as part of the full sample in Fig 2 . For each object there are two cr-split $`I`$-band exposures of unequal length available, together giving a typical exposure time of $`1800`$ seconds. The initial processing of the images, flat-fielding and bias removal, was carried out by the standard HST pipeline. The two exposures were then combined using the iraf task crrej, which successfully removes cosmic ray events using a sophisticated sigma-clipping algorithm. The next step in the reduction process was the fitting of a plane to the image with the 3CR galaxy masked out, to accurately determine the sky background while allowing for any residual flat-fielding gradients that may have been present. The final step was the production of a two-dimensional mask for each source which eliminated any companion objects, or aligned emission, from the model-fitting process, as well as any regions of the image that could have been biased by scattered light from nearby bright stars. In the vast majority of cases the high resolution provided by the HST allowed the identification, and subsequent masking, of companion objects and diffuse aligned emission in a straightforward fashion. For a small number of sources (3C252 & 3C239) the question of which areas of the images should be masked was more ambiguous, due to the HST images revealing double nuclei which are not apparent in near-infrared images (BLR 1997). As mentioned by BLR, the detection of two nuclear components in these sources suggests either that they are the products of recent mergers, or alternatively, contain a central dust lane. Given the undisturbed appearance of the K–band images of these sources (BLR 1997), it was decided that no attempt would be made to mask the central regions of the HST images before modelling. None of the objects comprising the 10-object sub-sample (see Section 4) appear to be highly disturbed in the HST images, and consequently return model fits which are largely insensitive to the details of aligned-emission masking. The advantage of being able to mask areas of the HST images, while still running the model fitting in two-dimensions, has allowed us to take greater advantage of the HST data than was possible with the one-dimensional analysis techniques employed by BLR. ### 2.1 Empirical PSF Determination Simple radial surface-brightness plots for the 16 objects chosen for modelling reveals that there is sufficient signal-to-noise to allow fitting to a typical radius of $`4^{\prime \prime }`$. It is therefore necessary to have a point spread function (PSF) to convolve our model galaxies with which has good signal-to-noise out to at least this radius. Due to the synthetic PSF’s produced by the tinytim software package (Krist 1995) being unable to reproduce the WFPC2’s scattered light halo outside a radius of $`1.5^{\prime \prime }`$, it is clear that an empirical PSF is required for the modelling of these data. A glance at Table 1 reveals that all of the exposures utilised here were imaged through either the F785LP or F814W filters. This required the acquisition of two relatively deep PSF’s imaged with the correct filter/chip combination, which were also located close to the average chip position of the 3CR galaxies, in order to avoid the noticeable positional variation of the WFPC2 PSF. An interrogation of the HST PSF search tool <sup>2</sup><sup>2</sup>2http://www.stsci.edu/instruments/wfpc2/ produced disappointing results, with all available PSF’s either too faint or too small in angular extent. Fortunately, two suitable stars were present on the exposures of 3C41 (F785LP) and 3C239 (F814W). Due to the need for sufficient depth in the PSF wings, both of these stars had saturated cores in even the shortest exposures. To overcome this problem a modified version of the PSF re-sampling technique described in McLure et al 1999b was implemented. Both PSF’s were only saturated within a radius of $`0.3^{\prime \prime }`$ of their core, well inside the radius where tinytim can accurately reproduce the empirical PSF. Making use of this, the two PSF’s had their core replaced with the equivalent tinytim model, the relative scaling being determined by matching the flux in an annulus between $`0.4^{\prime \prime }<r<0.7^{\prime \prime }`$. Another advantage of this approach is that due to tinytim’s ability to produce model PSF’s at up to 50 times oversampling, sub-pixel centring of the PSF can be matched to that of the galaxies to an accuracy of $`0.005^{\prime \prime }`$. Although not as crucial as with the quasar host work it is still important to overcome the severe undersampling of WFPC2, especially as several of the 3CR galaxies modelled proved to have a substantial point-source contribution. ## 3 Modelling The modelling of these objects proceeded in an identical fashion to that of the $`z0.2`$ AGN sample studied by McLure et al. (1999a) and Dunlop et al. (1999). Each object was modelled with both a standard Freeman disc template and de Vaucouleurs $`r^{1/4}`$ profile, with no a priori assumptions being made about position angle, axial ratio or any possible central point-source contribution. Prior to the modelling of each source a maximum radius, outside of which pixels would be excluded for insufficient signal-to-noise, was determined by an examination of the one-dimensional surface-brightness profile. The median value for this outer radius was $`4^{\prime \prime }`$, corresponding to a physical radius of $`33`$ kpc at the median redshift of $`z=0.8`$ ($`H_0=50,\mathrm{\Omega }_0=1`$). This is reasonably well matched to the outer radius of $`12^{\prime \prime }`$ ($`50`$ kpc) used for each of the radio galaxies in the $`z=0.2`$ AGN sample, ensuring that the same physical region of the galaxy light distribution is under investigation in both redshift regimes. The modelling technique strongly preferred a $`r^{1/4}`$ host galaxy for all $`16`$ objects analysed. The quality of the fits was good, with no significant residual flux. ## 4 Results In this section the scalelength, absolute luminosity and Kormendy relation results from our two-dimensional modelling of the $`z0.8`$ objects are presented. Due to the significant effect that choice of cosmology can have on angular diameter, cosmological dimming and look-back time over this redshift range, we give these results for a range of possible cosmologies. Four different representative scenarios are considered featuring two values of $`H_0`$ (50,70), together with both open ($`\mathrm{\Omega }_0=0.1`$) and flat ($`\mathrm{\Omega }_0=1`$) geometry. The results from the modelling of all 16 HST objects are presented in Table 1. The main conclusions reached from this modelling work are based largely on the results for a 10-object sub-sample. The sub-sample consists of the low-redshift end of the BLR sample and is illustrated in Fig 2. The remaining 6 objects ($`z0.9`$) have been excluded from the following analysis due to the incursion of the $`4000`$Å break into the HST F814W filter. The throughput of the F814W filter is shown in Fig 3 complete with the system response and CCD quantum efficiency. As can be seen from this Figure, for objects with redshifts $`z0.9`$ this filter will bridge the $`4000`$Å break, meaning that it can no longer be assumed that the detected flux has originated from the dominant old stellar population. Given that the alignment effect in radio galaxies is stronger in the rest-frame UV than at longer wavelengths it is to be expected that these objects are contaminated by a significant aligned component. The $`10`$-object sub-sample has the advantage of having a tight redshift distribution with a mean of $`0.74\pm 0.07`$ and a median of $`0.76`$, which is well matched by the distribution of the $`10`$ radio galaxies studied at low-$`z`$, with its mean redshift of $`0.20\pm 0.04`$ and median of $`0.20`$. This allows a direct comparison between the two samples without the added complication of allowing for significant evolutionary effects within the samples themselves. ### 4.1 Scalelengths The results of our determination of galaxy half-light radius ($`r_e`$) for the 10-objects in the $`z0.8`$ sub-sample are presented in Table 3. Two features of this table are immediately obvious. Firstly, it is clear that regardless of cosmology the derived half-light radii are remarkably consistent, with both the mean and median values agreeing to within 4 kpc. Secondly, the mean scalelength is systematically larger than the median in all four cosmologies, consistent with the suggestion from Fig 4 that the distribution of scalelengths from the full 16-object sample, and indeed our 20 radio-loud AGN at $`z0.2`$, has a substantial tail toward high values. Given that the scalelength results obtained by BLR are quoted assumming $`H_0=50,\mathrm{\Omega }_0=1.0`$ it is straightforward to investigate what differences exist between their scalelength determinations and those presented here. Due to the fact that most of the BLR scalelength information is derived from their K–band observations rather than the HST images, it is not the case that they have published a scalelength value for each of the objects re-modelled in this paper. In order to perform a comparison between the two sets of results it has been necessary to simply base the BLR figures on the 8-objects from our 10-object sub-sample for which they have derived a scalelength value. Assuming $`H_0=50,\mathrm{\Omega }_0=1.0`$ the mean scalelength of the 10-object sample is $`r_e=12.61\pm 2.26`$ kpc with a median of $`9.70`$ kpc. The corresponding BLR-derived values for $`8`$ of these $`10`$ objects are $`r_e=15.15\pm 2.71`$ kpc with a median of $`14.95`$ kpc. Given the small number statistics that are available, and the inherent difficulty in constraining galaxy scalelengths, the median is the more robust measure of the typical scalelength. Using the median it can be seen from these figures that the one-dimensional analysis technique employed by BLR has systematically overestimated the characteristic size of the 3CR galaxies by $`50\%`$. In order to check that the use of these two cut-down samples was not overtly biasing the results, a comparison was also performed between the full $`16`$-object sample and the $`12`$-objects from this sample for which there is also a BLR-derived scalelength figure available. The mean scalelength of the $`16`$-object sample is $`r_e=11.48\pm 1.59`$ kpc with a median of $`9.20`$ kpc. The corresponding BLR-derived values for $`12`$ of these objects are $`r_e=15.66\pm 1.91`$ kpc with a median of $`15.45`$ kpc. It can be seen from this that the use of the expanded samples strengthens the conclusion that the BLR scalelengths are overestimated, with the median BLR scalelength being $`70\%`$ greater than the two-dimensional modelling results presented here. ### 4.2 Absolute Magnitudes The absolute Cousins $`I`$-band magnitudes for the $`z0.8`$ and $`z0.2`$ sub-samples are listed in Table 4 for the four different cosmologies. The values shown are calculated from integrating the best-fit de Vaucouleurs profile to infinite radius in order to be consistent with the published results for the low-$`z`$ AGN hosts (McLure et al 1999a, Dunlop et al 1999). As was pointed out in Section 2 the objects comprising the sub-sample were imaged through two separate HST filters; F814W and F785LP. The F814W filter is a member of the standard HST filter set and closely mimics the Cousins $`I`$-band filter. As a result, it was decided to convert all of the integrated apparent magnitudes to their equivalent Cousins $`I`$-band value. The $`5`$ objects from the sub-sample imaged with the F814W filter had no correction applied to them since the similarity of the F814W and Cousins $`I`$-band filters is such that for elliptical galaxies the difference in magnitude is expected to be less than $`0.05`$ in all cases. This level of photometric accuracy is substantially greater than is possible in the face of the uncertainties in determining the host magnitudes. The throughput for the F785LP filter is significantly different from that of the Cousins $`I`$-band filter. The conversion from magnitudes obtained through this filter to Cousins $`I`$-band magnitudes is therefore more complicated. The original strategy to overcome this problem was to make use of the one object from the $`16`$-object sample (3C239) for which there are exposures in both filters available with comparable signal-to-noise. However, this object has the highest redshift, $`z=1.781`$, of all of the $`28`$ objects in the BLR sample and subsequently the two images were too faint to get a reliable conversion. In addition to this, 3C239 has a highly distorted morphology in the HST images (BLR 1997), making it unclear whether a conversion factor obtained from this object would be applicable to the mostly undistorted objects in the sub-sample. In light of this, the method used to convert the F785LP fluxes for each host to the equivalent F814W figure, and hence Cousins I-band magnitude, was to predict the count rates $`R_{object}`$ ($`e^1s^1`$pixel<sup>-1</sup>) of a source of apparent visual magnitude $`V`$ in the two filters using the filter ratios of Holtzman et al 1995. The cosmology-independent k-corrections for each object, produced by the blueward shifting of the $`I`$-band filter along the galaxy spectrum with increasing redshift, were calculated from the figures presented for a burst-elliptical galaxy by Rocca-Volmerange & Guiderdoni (1988). Given that the predictions produced by different spectral synthesis codes can differ significantly (Charlot et al 1996) it was considered worthwhile to make an independent check of the validity of these k-corrections. Due to the restriction of this analysis to the $`10`$-object sub-sample, for which the images sample galaxy light longward of the $`4000`$Å break, it is possible to estimate the necessary cosmological k-corrections by modelling the galaxy spectrum as a power-law of the form $`f_\nu \alpha \nu ^\alpha `$, where $`\alpha `$ is the spectral index. The value of $`\alpha `$ appropriate for radio galaxies imaged in the $`I`$-band was estimated to be $`\alpha 2`$ from a typical old ($`12`$Gyr) elliptical galaxy spectrum. Alternatively, the mean $`IJ`$ colour of the sub-sample objects can be used to estimate the spectral index, a process which also results in a figure of $`\alpha 2`$. Both of these estimates are in good agreement with the value of $`\alpha =1.82`$ needed to reproduce the k-corrections of Rocca-Volmerange and Guiderdoni (1988). The question of whether the absolute magnitudes presented in Table 4 are substantially different from the results previously obtained for the $`z=0.2`$ radio galaxies is addressed within Section 5. ### 4.3 The Kormendy Relation Given that the hosts of the $`10`$ radio galaxies within the sub-sample are well fitted by a standard de Vaucouleurs galaxy template, it is interesting to see whether the parameters obtained from these fits produce a Kormendy relation comparable to that followed by both low-$`z`$ inactive ellipticals (and the hosts of the $`z=0.2`$ AGN, see Section 5). The $`\mu _{1/2}`$ values required to construct the Kormendy relation have been corrected for the cosmological dimming of surface-brightness according to: $$\frac{I_1}{I_2}=\frac{(1+z_2)^{3+\alpha }}{(1+z_1)^{3+\alpha }}$$ (1) where a value of $`\alpha =1.8`$ has been assumed. The resulting $`\mu _{1/2}r_{1/2}`$ relations for the four cosmologies are given in Table 5. It is clear from this that the choice of cosmology makes no significant difference to the slope of the Kormendy relation (as expected given the small redshift range of the objects), and that in all cases this slope is consistent with that of $`3`$ displayed by inactive low-$`z`$ ellipticals (Kormendy 1977). It is worth noting that the $`r_{1/2}`$ and $`\mu _{1/2}`$ parameters from the one-dimensional modelling of BLR failed to produce a $`\mu _{1/2}r_{1/2}`$ relation which was consistent with the expected slope of $`3`$. The modelling results of BLR had the $`z1`$ 3CR galaxies lying along a constant luminosity slope of $`5`$, exactly as is expected when the galaxy luminosities have been well determined but the scalelengths have not been constrained (Abraham et al. 1992). The fact that the relation presented above is consistent with the Kormendy relation, within the errors, can therefore be taken as further evidence that our two-dimensional modelling has been much more successful in constraining the scalelengths of the sub-sample objects. The possibility of using the Kormendy relations derived above to test for the effects of passive or dynamical evolution is explored in section 5.3. ## 5 Comparison with low-redshift Radio Galaxies ### 5.1 Scalelengths The derived median and mean scalelengths for the 10-source $`z0.2`$ radio galaxy sample from our low-$`z`$ AGN host galaxy study (McLure et al. 1999a, Dunlop et al. 1999) are presented in Table 6. A comparison of these results with those presented for the $`z0.8`$ sub-sample in Table 3 shows the two groups of galaxies to have very similar characteristic scalelengths. The poorest agreement between the two sets of results is for the $`\mathrm{\Omega }=1.0,H_0=50`$ cosmology. However, even in this case the figures are consistent to within the errors. In the other three cosmologies the median and mean scalelengths for the two samples are basically identical, differing by $`0.5`$ kpc in all cases. The similarity between the scalelength distributions of the two sub-samples is confirmed by an application of the Kolmogorov–Smirnov (KS) test. For the two $`\mathrm{\Omega }_0=1.0`$ cosmologies the KS test returns a probability of 0.68 that the two samples are drawn from the same underlying distribution. This conclusion is even stronger in the $`\mathrm{\Omega }_0=0.1`$ cosmologies where the KS test returns a probability of 0.97 that the two distributions are the same. Therefore, we find no evidence that the galaxies comprising the $`z0.8`$ sub-sample are systematically larger than their $`z0.2`$ counterparts. ### 5.2 Absolute Luminosity In order to allow a direct comparison of the characteristic luminosities of the low- and high-redshift radio galaxy sub-samples, and thereby look for any evidence of significant merger activity, it is first necessary to correct for the different filters and the expected passive evolution of the stellar populations. To achieve this it was decided to convert the $`z0.2`$ $`R`$-band magnitudes to their $`I`$-band equivalents, and then to predict the brightening of the stellar population between $`z=0.2`$ and $`z=0.8`$ due to passive evolution alone. The present-day colour of an elliptical galaxy formed at high redshift ($`z3`$) has been taken as $`RI=0.7`$ (Fukugita et al. 1995). The corrections to be made for the effects of passive evolution have been calculated using the synthetic galaxy spectral models of Jimenez et al. (1996) using our chosen set of four possible cosmologies. Several different galaxy formation redshifts were considered to investigate their effects on the resulting correction. A comparison between the predicted amount of $`I`$-band passive evolution between $`z=0.2`$ and $`z=0.8`$ and the difference in median absolute luminosity of the two samples is presented in Table 7. It is immediately clear from these results that the difference in absolute $`I`$-band luminosity between the $`z0.2`$ and $`z0.8`$ samples is inconsistent with the amount of passive evolution predicted by the stellar synthesis models within the two Einstein-de Sitter cosmologies. This is still true even in the $`\mathrm{\Omega }_0=1`$ model which allows present-day ellipticals to be as old as possible ($`z_{for}=10,H_0=50`$); this model still requires 2.5 times more magnitudes of passive evolution than is seen in the data. Adoption of either of the open cosmologies results in a measured difference in absolute luminosity of $`0.5`$ mags between the two galaxy samples (columns 3 & 4). As can be seen from Table 7, this luminosity difference is formally consistent with the expected passive evolution in both open cosmologies, for all alternative formation redshifts. However, it is interesting to note from Table 7 that the $`H_0=70,\mathrm{\Omega }_0=0.1`$ cosmology clearly provides the best match between the data and the spectrophotometric model predictions. In this cosmological picture the model predictions are perfectly consistent with the data for all star formation redshifts of $`z4`$, in good agreement with recent discoveries of old stellar population ellipticals at redshifts of $`z=1.52`$ (Dunlop et al. 1996, Spinrad et al. 1997, Stiavelli et al. 1999 ). ### 5.3 Kormendy Relation If the two samples of radio galaxies can truly be linked by a single population of passively evolving ellipticals it is to be expected that they should follow Kormendy relations which are identical except for a simple vertical shift in surface-brightness. After making the appropriate surface-brightness corrections (Section 4.3) and $`RI`$-band filter transformation (Section 5.2) the least-squares fit to the Kormendy relations formed by the $`z0.2`$ radio galaxies are shown in Table 8. It can be seen from this that, as for the $`z0.8`$ galaxies, the choice of cosmology makes little difference to the slope of the Kormendy relation, although the normalization does change significantly. It is also apparent that the low-$`z`$ objects appear to follow a substantially flatter relation, with a slope of 2.9 instead of 3.5, although clearly both values are consistent with the expected slope $`3`$ (due to the fairly substantial formal error in the fitted slope, which is predominantly a result of the small sample size and lack of dynamic range in radius). To facilitate a fair comparison of the high- and low-$`z`$ radio galaxy Kormendy relations, the least-squares fitting was repeated with an enforced intermediate slope of 3.20 . If it is indeed the case that these two galaxy populations can be linked by passive evolution alone then the vertical shift required to reconcile the Kormendy relations should be in good agreement with both that required to match the absolute magnitudes, and the passive evolution predictions of the stellar synthesis modelling. To explore what range of vertical magnitude shifts are allowed by the data, we performed the least-squares fitting (with fixed slope=3.20) not only with the best estimate of $`r_e`$ for each galaxy, but with the scalelengths of the $`z0.8`$ sub-sample shortened and lengthened such that the KS test showed its scalelength distribution to differ from that of the $`z0.2`$ sub-sample at the $`1\sigma `$ level. The results of this process are presented in Table 9. A comparison of the figures from Table 9 with those of Table 7 shows that in the $`\mathrm{\Omega }_0=1.0`$ models considered here it is only possible to reconcile the luminosity evolution predicted by the Kormendy relations with that predicted by the spectral modelling and the absolute magnitude distributions, separately. In these cosmologies it is impossible to force both alternative measures of the amount of luminosity evolution, and the predictions of the spectral modelling, into agreement. In contrast, the $`\pm 1\sigma `$ predictions from the Kormendy relations in the $`\mathrm{\Omega }_0=0.1`$ cosmologies comfortably bracket the offsets given in Table 7, thus leaving both alternative measures of the luminosity evolution in excellent agreement. In theory, the Kormendy relations for the low and high-z sub-samples presented here offer an opportunity to constrain both cosmology and the prevalence of merger activity. However, in practise the effects of these are very closely coupled. For example, our results can be reproduced either by pure passive evolution since $`z0.8`$ in an open Universe, or by modest growth ($`20\%`$ growth in scalelength & luminosity) since $`z0.8`$ in an Einstein-de Sitter Universe. Despite this inherent degeneracy the Kormendy relations can firmly exclude very strong growth ($`50\%`$) from mergers in the redshift range $`0.2<z<0.8`$. ## 6 Conclusion The results from a thorough re-examination of the BLR HST images of a sample of $`z1`$ 3CR radio galaxies have been presented. It has been shown that, in terms of scalelength, absolute magnitudes and Kormendy relation there are no significant differences between $`z0.8`$ and $`z0.2`$ 3CR radio galaxies. The two populations appear to be fully consistent with being comprised of old stellar populations formed at high redshift and evolving passively thereafter. It is obviously true that the fact that both the high and low-redshift radio galaxies have absolute luminosities consistent with pure passive evolution can easily be reconciled with the involvement of some dynamic evolution. In the dynamical model this simply requires that the proto-galactic clumps which merge to produce the final galaxies where formed reasonably coevally. The results presented here do not then require that these radio galaxies must have formed in a monolithic collapse at a single redshift. However, they do suggest strongly that the vast majority of merger activity within this population of massive ellipticals must have been completed before $`z1`$. Although semi-analytical galaxy formation models in a $`\mathrm{\Omega }_0=0.3,\mathrm{\Lambda }=0.7`$ cosmology predict as much as 70% of present day ellipticals to already be in place by $`z1`$ (Kauffman & Charlot 1999), it is worth remembering that the same is not true of BCGs, which are still predicted to grow by a factor of $`24`$ between $`z=1`$ and the present day (Aragón-Salamanca et al. 1998). The crucial question still to be answered is whether or not it can be proven that a significant difference in cluster environment does exist between the two populations. If it can be shown that this is definitely the case then the argument forwarded by BLR that we are observing the effects of dynamical evolution producing the characteristic mass required for powerful radio emission at the two different epochs remains tenable; the correlation between luminosity and scalelength would actually predict the $`z0.8`$ and $`z0.2`$ sub-samples to have similar scalelength distributions. However, as was discussed in Section 1, while there have been numerous surveys carried out recently tackling the environments of high-$`z`$ radio sources, the work on the low-$`z`$ environments looks to be subject to possible systematic error. This suspicion is further strengthened by our new deep R–band HST images of the 10 $`z0.2`$ radio galaxy sub-sample (McLure et al. 1999a, Dunlop et al. 2000). These images show numbers of apparent companion objects which appear consistent with Abell classes $`02`$. This is in good agreement with our recent work to model the first to fourth-ranked brightest cluster galaxies (BCG) in clusters spanning Abell classes $`14`$, imaged with HST (Dunlop et al. 1999). This work has shown that the scalelengths of these objects are not significantly different from the radio galaxy scalelength results presented here. In an attempt to disentangle the radio galaxy environment problem, more near-infrared observations of the low-$`z`$ radio galaxies and quasars from our HST programme are planned, to compliment our existing R and B–band images of this sample. This combined dataset will allow a fully qualitative investigation of the cluster environments of the low-$`z`$ radio galaxies with which to compare with the existing high-$`z`$ environmental data. If it transpires that there is no significant difference in 3CR environments at high and low-$`z`$, then a much simpler picture of powerful radio galaxy evolution emerges. Combined with the scalelength evidence presented here, this would suggest that the host galaxies of powerful radio sources are basically the same sort of objects at all redshifts from $`z=01`$. In this picture the fall-off in 3CR radio power over this redshift range would simply be due to the progressively smaller amount of gas available to feed the central engine. ## 7 Acknowledgements We thank the referee Philip Best for comments which undoubtedly improved a number of aspects of this paper. Based on observations with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract No. NAS5-26555. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. MJK acknowledges the award of a PPARC PDRA, and also acknowledges support for this work provided by NASA through grant numbers O0548 and O0573 from the Space Telescope Science Institute, which is operated by AURA, Inc., under NASA contract NAS5-26555. RJM acknowledges a PPARC studentship. ## 8 References Abraham R.G., Crawford C.S., McHardy I.M., 1992, ApJ, 401, 474 Aragón-Salamanca A., Baugh C.M., Kauffmann G., 1998, MNRAS, 292, 427 Best P.N., Longair M.S., Röttgering H.J.A., 1997, MNRAS, 292, 758 Best P.N., Longair M.S., Röttgering H.J.A., 1998, MNRAS, 295, 549 Bruzual G., Charlot S., 1993, ApJ, 405, 538 Charlot, S., Worthey G., Bressan. A., 1996, ApJ, 457, 43 Collins C.A., Mann R.G., 1998, MNRAS, 297, 128 Dickinson M., 1997, in Tanvir N.R., Aragon-Salamanca A., Wall J.V., eds, HST and the high redshift Universe. Singapore: World Scientific, p.207 Dunlop J.S., Peacock J.A., 1993, MNRAS, 262, 936 Dunlop J.S., Taylor G.L., Hughes D.H., Robson E.I., 1993, MNRAS, 264, 455 Dunlop J.S., et al., 1996, Nature, 381, 581 Dunlop J.S., et al., 2000, MNRAS, in preparation Eales S., Rawlings S., Law-Green D., Cotter G., Lacy M., 1997, MNRAS, 291, 593 Fukugita M., Shimasaku K., Ichikawa T., 1995, PASP, 107, 945 Hill G.J., Lilly S.J., 1991, ApJ, 367, 1 Holtzman J.A., et al, 1995, PASP, 107, 1065 Jimenez R., MacDonald J., 1996, MNRAS, 283, 721 Kauffmann G., Charlot S., 1999, astro-ph/9810031 Kormendy J., 1977, ApJ, 217, 406 Krist J., 1995, Astronomical Data Analysis and Systems IV, ASP Conference Series, Vol 77, R.A. Shaw, H.E. Payne, and J.J.E. Hayes, eds, 349 Laing R.A., Riley J.M., Longair M.S., 1983, MNRAS, 204, 151 Leyshon G., Eales S.A., 1998, MNRAS, 295,10L Lilly S.J., Longair M.S., 1984, MNRAS, 211, 833 Lilly S.J., Prestage R.M., 1987, MNRAS, 225, 531 Oke J.B., Gunn J.E., 1983, ApJ, 226, 713 Madore B.F., et al., 1999, ApJ, 515, 29 McCarthy P.J., van Breugel W.J.M., Spinrad H., Djorgovski S., 1987, ApJ, 321, L29 McLure R.J., et al., 1999a, MNRAS, 308, 377 McLure R.J., et al., 1999b, MNRAS, submitted Perlmutter S., et al., 1999, ApJ, 517, 565 Prestage R.M., Peacock J.A., 1988, MNRAS, 230, 131 Rocca-Volmerange B., Guiderdoni B., 1988, A&AS, 75, 93 Schombert J.M., 1987, ApJS, 64, 643 Spinrad H., et al., 1997, ApJ, 484, 581 Stiavelli M., et al., 1999, A&A, 343, L25 Taylor G.T., Dunlop J.S., Hughes D.H., Robson E.I., 1996, MNRAS, 283, 930
no-problem/9908/quant-ph9908053.html
ar5iv
text
# MR spectroscopy with parabolic magnetic field: spin-oscillator coupling effect ## Abstract The spectrum of the spin particle in oscillatory potential subjected to external parabolic magnetic field $`𝐁=(B_0+Gx+\overline{G}x^2)\widehat{𝐳}`$ is obtained. The structure of energy levels of the considered system allows to identify the frequency of the oscillator via the spectrum of spin sublevels coming only from one oscillatory level. The effect is due to the gradient terms in the form of the field. In a recent decade one can observe a rapid grow of interest in the Stern-Gerlach (SG) interaction within MR spectroscopy . The main reasons are the following: (i) partial saturation of the possibilities of the conventional MR methods based on Bloch-Purcell paradigm (low sensitivity, especially for the nuclear MR (NMR), growing costs of aparature, existence of cheaper, competitive methods like e.g. ultrasonic ones), (ii) recent reports on new and promising effects (e.g. atomic force microscopy , MR microimaging , MR spectroscopy for selected Zeeman states (iii) the need for new solutions in the domain of NMR quantum computing (the main problem in the recent implementations of quantum computing is low signal intensity) (iv) the recent development in the domain of ferromagnetics and superconductors allowing to produce strong magnetic fields of required shapes. In this situation one is seeking for new, both theoretical and experimental solutions for MR, in particular by exploiting the SG interaction. Recently Sidles has proposed the interesting model of spin-$`\frac{1}{2}`$ particle mounted to the harmonic oscillator. The particle is assumed to interact with the magnetic field via Stern-Gerlach Hamiltonian plus some time-dependent, constant in space magnetic field. He solved the model and showed that the wavepacket of the particle exhibits the splitting into two parts moving in opposite directions. He also predicted the importance of possible oscillator based NMR spectroscopy for molecular imaging devices. In this Letter we consider the coupled spin-oscillator system in external heterogeneous magnetic field. The model is solved exactly and provides physical spectrum in which the oscillator frequency is coupled to the field derivatives parameters. It implies that in principle one can recognise the frequency of oscillator basing only on signals coming form transitions between sublevels of one oscillatory level. Let us consider the particle that subjects to the simple dynamic due to the harmonic oscillator Hamiltonian $`_{osc}^\mathrm{\Omega }(xa)`$. Here $`a`$ is the position of the minimum of the oscillator potential with respect to the origin of the reference frame (we imagine that the particle is in a sample of finite dimension $`l`$ situated in the origin of the reference frame, so that $`a`$ can vary within the region $`(l,l)`$). The Hamiltonian of the particle $$_{osc}^\mathrm{\Omega }(x)=\frac{\mathrm{}^2}{2m}\frac{d^2}{dx^2}+\frac{m\mathrm{\Omega }^2x^2}{2}.$$ (1) has the following spectrum and corresponding eigenvectors: $`_n=\mathrm{}\mathrm{\Omega }(n+{\displaystyle \frac{1}{2}})`$ (2) $`\mathrm{\Psi }_n(\mathrm{\Omega };x)=({\displaystyle \frac{m\mathrm{\Omega }}{2^{2n}(n!)^2\mathrm{}\pi }})^{\frac{1}{4}}H_n(\sqrt{{\displaystyle \frac{m\mathrm{\Omega }}{\mathrm{}}}}x)exp({\displaystyle \frac{m\mathrm{\Omega }}{2\mathrm{}}}x^2)`$ (3) with Hermite polynomials $`H_n`$. Suppose now that the particle has the spin $`S`$ and recall that the quantum operator corresponding to the projection of the spin onto the $`z`$ axis is described as $`\widehat{\text{I}}=_{M=S}^S\mathrm{}M|MM|`$ with the corresponding spectrum and eigenvectors $`\mathrm{}M,|M`$. Consider the inhomogeneous magnetic field $$\text{B}(x)B(x)\widehat{\text{z}}=(B_0+Gx+\overline{G}x^2)\widehat{\text{z}}$$ (4) along $`\widehat{\text{z}}`$ axis. Here $`G`$, $`\overline{G}`$ stands for the gradient and second derivative parameters. Note that the field has the value $`\text{B}_0=B_0\widehat{\text{z}}`$ for $`x=0`$ an that in general it depends quadratically on the spatial coordinate $`x`$. Now if we put our particle in the field (4) then its Hamiltonian becomes $$=_{osc}^\mathrm{\Omega }(xa)I\gamma B(x)\widehat{\text{z}}\widehat{\text{I}}$$ (5) In the above formula the difference between spatial and spin degrees of freedom has been stressed. In particular the identity operator $`I`$ has been used. The new energy spectrum of our particle can be found seeking the eigenvectors in the form $`|\varphi |M`$ ($`M=S,S+1,\mathrm{},S`$). This leads effectively to the sequence of $`2S+1`$ $`M`$-dependent shifted quantum oscillators. Eigenvalues each of those oscillators can be easily solved and the final energy spectrum of the particle is given by $`E_{M,n}=\mathrm{}\sqrt{\mathrm{\Omega }^2{\displaystyle \frac{2\gamma \overline{G}\mathrm{}M}{m}}}(n+{\displaystyle \frac{1}{2}})\gamma (B_0+Ga+\overline{G}a^2)\mathrm{}M`$ (6) $`{\displaystyle \frac{\gamma ^2(G+2\overline{G}a)^2\mathrm{}^2M^2}{2m(\mathrm{\Omega }^2\frac{2\gamma \overline{G}\mathrm{}M}{m})}}`$ (7) (8) and the corresponding eigenvectors can be written in the form $`\mathrm{\Phi }_{M,n}=|\varphi _{M,n}|M`$ (9) The spatial coordinate function is defined by the oscillator eigenvectors (3) in the following way $$\varphi _{M,n}(x)\mathrm{\Psi }_n(\stackrel{~}{\mathrm{\Omega }}_M;xa\frac{\gamma (G+2\overline{G}a)\mathrm{}^2M}{m\stackrel{~}{\mathrm{\Omega }}_M^2})$$ (10) where $`\stackrel{~}{\mathrm{\Omega }}_M\sqrt{\mathrm{\Omega }^2\frac{2\gamma \overline{G}\mathrm{}M}{m}}`$. It is important to note that the squared component of the field represented by its second derivative $`\overline{G}`$ must not be too large as then it suppresses the squared term of the particle oscillator potential $`\frac{m\mathrm{\Omega }^2q^2}{2}`$ leading to Hamiltonian unbounded form below. This occurs for the values violating the inequality $$\overline{G}<\frac{m\mathrm{\Omega }^2}{2\gamma \mathrm{}M_{max}}.$$ (11) where $`M_{max}=S`$. In such case the first term of the spectrum (8) becomes imaginary for some $`M`$. Now, if the particle were a light, spin endowed part of molecule (the model which we shall discuss subsequently), this could be interpreted as the dissociation of the molecule caused by the strong gradient of the magnetic field (herewith, we will keep, a bit roughly, the term “dissociation”). The last inequality establishes the boundaries of our model. In this context it is convenient to introduce the new discrete parameter defined by $$\overline{M}=\frac{2\gamma \overline{G}\mathrm{}M}{\mathrm{\Omega }^2m}$$ (12) Using $`\overline{M}`$ the inequality (11) writes as $$\overline{M}_{max}<1$$ (13) i.e. the dissociation occurs if $`\overline{M}`$ reaches the unity. In terms of $`\overline{M}`$, the dissociation occurs if $`\overline{M}`$ reaches unity, so that the scope of the presented model is described by inequality $`\overline{M}<1`$. Roughly speaking $`\overline{M}`$ is equal to the ratio of $`M`$ to the number of the maximal spin level for which dissociation occurs. As we are interested in the effects caused by gradients, we put $`B_0=0`$. Then the energy spectrum can be written in the following form $`E_{M,n}=\mathrm{}\mathrm{\Omega }({\displaystyle \frac{1}{2}}+n)\sqrt{1\overline{M}}{\displaystyle \frac{m\mathrm{\Omega }^2(G/\overline{G})a}{2}}{\displaystyle \frac{\overline{M}}{1\overline{M}}}`$ (14) $`{\displaystyle \frac{m\mathrm{\Omega }^2a^2}{2}}{\displaystyle \frac{\overline{M}}{1\overline{M}}}{\displaystyle \frac{m\mathrm{\Omega }^2(G/\overline{G})^2}{2}}{\displaystyle \frac{\overline{M}^2}{4(1\overline{M})}}`$ (15) Consider first the case of infinitely small sample concentrated in the origin of the reference frame (i.e. put $`a=0`$). Then we obtain especially appealing form of energy spectrum $$E_{M,n}(a=0)=E_{osc}^{qu}(\mathrm{\Omega },n)\sqrt{1\overline{M}}E_{osc}^{cl}(\mathrm{\Omega },G/\overline{G})\frac{\overline{M}^2}{4(1\overline{M})},$$ (16) where $`E_{osc}^{qu}=\mathrm{}\mathrm{\Omega }^2(\frac{1}{2}+n)`$ and $`E_{osc}^{cl}(\mathrm{\Omega },G/\overline{G})=\frac{m\mathrm{\Omega }^2(G/\overline{G})^2}{2}`$. Here we have two competitive terms: the first one represents the energy of quantum oscillator of frequency $`\mathrm{\Omega }`$ (i.e. the energy of the particle in absence of the field) while the second one is nothing but the total energy of the classical oscillator of the same frequency, of the amplitude determined by the shape of the field. Both terms have weights depending on the scaled spin number. Let us now keep the value of $`\overline{G}`$ constant (suitably chosen, in order not to be too close to the dissociation regime, e.g. to have $`\overline{M}0.1`$). Now varying the linear gradient $`G`$ we obtain smooth transition from quantum oscillator regime to the classical one. Provided $`G`$ is sufficiently large, so that the classical part dominates, we have a kind of amplification of the vibrations of the particle by the magnetic field. In both cases the frequency $`\mathrm{\Omega }`$ determines the ratio of splitting of the spin levels. Thus by measuring the latter, we can obtain the information about the value of the frequency. Note that, remarkably, the same is impossible in homogeneous field: the possibility of monitoring vibrational frequency via magnetic resonance is exclusively due to the gradient terms. Note that the double “quantum-classical oscillator” structure of the energy levels is due to quadratic gradient term. Indeed, for $`\overline{G}=0`$ we we cannot use the scaled spin level number $`\overline{M}`$ as the latter is defined for $`\overline{G}0`$. In such a case we have no ”dissociation” as there is no denominator singularity in formula (8). If we put, instead, $`G=0`$ (with $`a0`$ we still have two oscillators, however the latter one follows from finite dimensions of the sample (so that the centre of the oscillations may feel nonzero value of magnetic field) $$E_{M,n}(G=0)=E_{osc}^{qu}(\mathrm{\Omega },n)\sqrt{1\overline{M}}E_{osc}^{cl}(\mathrm{\Omega },a)\frac{\overline{M}^2}{1\overline{M}}$$ (17) On Fig. 1 we present these levels for electronic MR (EMR) versus quadratic gradient $`\overline{G}`$ for the system of spin $`S=3/2`$ (we keep $`G`$ being nonzero, but weak). One can observe remarkable crossing points for sufficiently large $`\overline{G}`$. Again, for a given value of $`\overline{G}`$ the level difference is determined by the vibrational frequency $`\mathrm{\Omega }`$. Let us summarise our results. The very important feature of the obtained spectrum of the particle is that the spin quantum number $`M`$ has been coupled to the characteristic frequency $`\mathrm{\Omega }`$ of the particle. This means that now for fixed quantum oscillator number $`n`$ the M-dependent energy sublevels form the structure which depends on the characteristic frequency $`\mathrm{\Omega }`$ of the particle. Indeed, putting $`G=\overline{G}=0`$ one can immediately see that it does not happen for uniform field $`\text{B}(x)=\text{B}_0`$. It is remarkable result for the following reasons. Imagine the molecules with spin-less heavy “core” and spin $`S`$ light part coupled to it. Assume that we have a mixture of unknown molecules of that kind, each of them possessing unknown characteristic frequencies $`\mathrm{\Omega }_1`$, $`\mathrm{\Omega }_2`$, … . Then the fact that the spectrum (8) (with nonzero $`G`$ or $`\overline{G}`$) depends on the frequency will provide the possibility to identify them basing directly on the NMR techniques that generically deal with $`2S+1`$ spin sublevels of the given level. It is interesting that if we have $`\overline{G}=0`$ and nonzero $`G`$ then the $`n`$ dependent term does not depend on quantum number $`M`$. Of course, the structure of spin sublevels does not depend on $`n`$. Then all the molecules of one kind (distinguished by $`\mathrm{\Omega }`$) would give the same signal from $`M`$ sublevel of all oscillatory levels. A very interesting feature of the considered interaction is the emerging amplification of the vibrational mode to the form of classical oscillator. The effect is due to nonzero $`\overline{G}`$. The amplitude is determined by the shape of the field, and the amplification is obtained for high values of the linear gradient. Concerning the experimental realiation of the presented spin-oscillator coupling effect note that for the typical molecular frequences ($`10^{14}Hz`$) one would need enormously strong gradients to observe the effect. Indeed, the strongest gradients obtained so far amount to $`10^6T/m`$ . Now, for molecule with incorporated atom of oxygene the above gradients allow to realize the effect for frequences up to $`10^4Hz`$. Besides the highest frequency for which the effect can be observed satisfies roughly $`\overline{G}\mathrm{\Omega }^2`$. On the other hand, there exist molecular systems that exhibit low frequences . Thus the effect is difficult but still possible to realize experimentally. In conclusion, we have demonstrated that MR spectroscopy with parabolic magnetic field provides subtle spin-oscillator coupling effects, that allow to select the characteristic frequency of the system. It seems to open new prospects for the MR applications such as quantum level separations or quantum computation . Finally, we expect that the idea of incorporating higher order gradients will prove fruitful in looking for new solutions within MR domain. We also believe that the obtained results will stimulate development of techniques of producing strong gradient field of different shapes. We are grateful to Prof. Zdzisław Paja̧k for very helpful discussion. Cz. L. was supported by the Polish Committee for Scientific Research, contract No. 2P03B 14812. M. H., P. H. and R.H. were partially supported by Polish Committee for Scientific Research, Contract No. 2 P03B 103 16 and the European Science Foundation. M.H. and P.H. also acknowledge the support from the Foundation for Polish Science.
no-problem/9908/astro-ph9908127.html
ar5iv
text
# The Distribution of High and Low Redshift Type Ia Supernovae ## 1 Introduction Statistical studies of SNe give us information about the Universe on all scales. They give us clues about the progenitors of SNe (impacting stellar evolution), reveal the galactic SN rate (influencing galaxy evolution), and are being used to determine cosmological parameters (the evolution of the Universe). For all of these tasks, obtaining an unbiased sample of SNe, and thus fully understanding the selection effects involved in SN discovery, is essential. This work will focus on selection effects that may affect the use of SNe Ia as distance indicators, so only SNe Ia are discussed. ### 1.1 SNe as Standard Candles All SNe Ia are thought to be caused by the thermonuclear explosions of carbon–oxygen (CO) white dwarfs in binary systems (Hoyle & Fowler 1960). Their low dispersion in absolute magnitude allows them to be used as “calibrated candles,” though some care must be exercised if they are to be used in this capacity. The first Hubble diagram constructed from Type I SNe by Kowal (1968) revealed a dispersion in photographic magnitude of $``$0.6 mag. Barbon et al. (1973) noted that Type I SNe were not a homogeneous class — some light curves declined faster than others. Pskovskii (1977,1984) then suggested that the light curve decline rate is correlated with the luminosity of SNe. Slower declining SNe Ia are intrinsically more luminous than those with fast-declining light curves. The separation of the Type Ib (Elias et al. 1985) and Ic subclasses (Wheeler & Harkness 1986), and the shift to modern CCD detectors allowed Phillips (1993) to establish a decline rate parameter, $`\mathrm{\Delta }m_{15}(B)`$, the decline in blue magnitude from maximum after 15 days, and to conclusively show that it was correlated with the absolute magnitudes of SNe at maximum. In a study of 29 SNe Ia, Hamuy et al. (1996) have shown that correcting the absolute magnitudes at maximum in this way reduces the dispersion from $`\sigma `$= 0.38, 0.26, 0.19 to $`\sigma `$= 0.17, 0.14, 0.13 in the B, V, and I bands respectively. Another approach, the Multicolor Light Curve Shape (MLCS) technique (Riess, Press, & Kirshner, 1996), makes use of the entire shape of the light curve in various colors to reduce the dispersion at maximum light to $`\sigma 0.15`$ magnitudes. These results have given promise that SNe Ia may be used as calibrated candles to determine large scale motions of the local group (Riess, Press, & Kirshner, 1995), and the cosmological parameters H<sub>0</sub>, $`\mathrm{\Omega }`$, and $`\mathrm{\Lambda }`$ (Goobar & Perlmutter 1995). Several searches have already discovered a combined number of $``$ 100 high redshift SNe ($`0.3z1.0`$) for this purpose (Riess et al. 1998, Perlmutter et al. 1999), with the remarkable suggestion that $`\mathrm{\Lambda }`$ is finite and positive. ### 1.2 The Shaw Effect Shaw (1979) first characterized a selection effect in supernova searches that had long been suspected. On photographic plates, supernovae are less frequently discovered in the often overexposed central regions of distant galaxies. Shaw estimated that $``$50% of supernovae are lost within the central 8 kpc of galaxies beyond 150 Mpc. The effect is reduced for closer galaxies. A correction factor must be introduced to account for this effect when determining SN rates. Bartunov et al. (1992) and Cappellaro et al. (1993, 1997) find that supernova searches experience the Shaw effect to varying degrees, so the correction factor for each search can be different. To quantify the Shaw effect, one must compare supernovae discovered photographically to an “unbiased” sample. This has been attempted with varying degrees of success, but so far each comparison group has had problems. Out of necessity, Shaw (1979), in his own words, made the “unwarranted” assumption that supernovae discovered in galaxies closer than 33 Mpc were free from this effect. Prior to this, the loss of SNe in the inner regions of galaxies was estimated by extrapolation of the SN rate in the outer regions of galaxies to a central peak following the light distribution (Johnson & MacLeod 1963; McCarthy, 1974; Barbon et al. 1975). Bartunov et al. (1992) confirmed Shaw’s suspicion that even the nearby sample is not free from bias. Bartunov et al. and Cappellaro et al. used combined data from visual and CCD searches as the control group. ### 1.3 Differences in SN Properties with Galactocentric Distance As pointed out by Wang, Höflich & Wheeler (1997, hereafter WHW), SN properties appear to vary with projected galactocentric distance (PGD). Using data from 40 local well-studied SNe Ia, WHW find that SNe located more than 7.5 kpc from the centers of galaxies show 3–4 times lower scatter in maximum brightness than those projected within that radius. (Note that the projected distance is the minimum distance the SN can be from the center of the galaxy. Many events at small PGD are actually located far from the center of the host galaxy.) A reduction of scatter with PGD is also apparent in the $`B_{\mathrm{max}}V_{\mathrm{max}}`$ colors of the SNe. In the absence of corrections for light curve shape, SNe at higher projected galactocentric distance are a more homogeneous group and should be better for use as distance indicators. The reddest and bluest, brightest and dimmest SNe are located near the galactic center, so extinction alone cannot explain the higher scatter in this region. Indeed, Riess et al. (1999) find that scatter still remains in these relationships when the SNe are corrected for extinction using the MLCS technique. The scatter is reduced after the SNe are corrected for the light curve decline relationship, leaving no apparent trend with PGD. As suggested by WHW and confirmed by Riess et al. (1999), SNe in the outer regions of galaxies show systematic differences in luminosity with respect to those with smaller projected separations from the host center. After correcting for extinction, Riess et al. find that SNe at distances of 10 kpc or more from the centers of their host galaxies are dimmer by about 0.3 mag than the mean of those (projected) inside 10 kpc. Riess et al. also note that this effect may be related to one pointed out by Hamuy et al. (1996) and Branch et al. (1996) — SNe in early-type hosts tend to be dimmer. At high PGD, the sample is dominated by SNe in elliptical host galaxies. Possible sources for the variation in SN properties with projected distance (if intrinsic) include metallicity gradients within the galaxy or differences in progenitor systems between the disks and bulges. ## 2 The sample Five groups of SNe Ia were studied in this work - the local sample discovered with CCDs, the local SNe discovered photographically, the high redshift SNe discovered by the Supernova Cosmology Project, the local SNe used to calibrate the brightness of the high–$`z`$ SNe, and eleven very local supernovae discovered in galaxies with Cepheid distances that are used to tie the SN distance scale to the Cepheid distance scale. The IAU Circular “List of Supernovae” web page<sup>1</sup><sup>1</sup>1http://cfa-www.harvard.edu/cfa/ps/lists/Supernovae.html was used to determine the initial sample. The Asiago Supernova Catalog (Barbon et al. 1989) was used to provide missing data and the recession velocities of the host galaxies. Any recession velocities not in the Asiago catalog were taken from the NASA/IPAC Extragalactic Database (NED)<sup>2</sup><sup>2</sup>2http://nedwww.ipac.caltech.edu/. The sample was divided into high and low redshift at $`z=0.3`$. The SNe were retained if they met the following criteria: * Type Ia * Projected offset of SN from center of host galaxy is known * Recession velocity or distance to host galaxy is known * Method of discovery is known Distances to galaxies with recession velocities greater than 3000 km s<sup>-1</sup> were computed using H<sub>0</sub>= 65 km s<sup>-1</sup> Mpc<sup>-1</sup>. For galaxies under 3000 km s<sup>-1</sup> distances were taken from Tully (1985) and scaled to H<sub>0</sub>= 65. The distances to high redshift SNe were calculated using the angular size distance for the $`\mathrm{\Lambda }=0`$ model: $$D=\frac{2c}{H_0}\left(\frac{(2\mathrm{\Omega }+\mathrm{\Omega }z)(2\mathrm{\Omega })\sqrt{1+\mathrm{\Omega }z}}{\mathrm{\Omega }^2(1+z)^2}\right)$$ (Fukugita et al. 1992) assuming $`H_0=65`$, $`\mathrm{\Omega }=0.2`$. The method of discovery (photographic, visual, CCD) was not listed in any database, so this information was obtained directly from the IAU Circulars reporting the discovery of each SN. When SNe are very near to the center of the host galaxy, positions are sometimes reported as “very close” to the center, and no separation in arcseconds is given. Four supernovae fell into this category, SNe 1994U, 1996am, 1997dg, and 1998ci. All are in the sample of local SNe Ia discovered with CCDs. Removing these SNe from the sample would have biased the results, so they were retained and assigned zero offset from the center of the galaxy. Presumably these SNe were located so close to the center of their host galaxy that the offset can be considered negligible. The high redshift SNe used were primarily those discovered by the Supernova Cosmology Project. Finding charts were downloaded from the group’s website<sup>3</sup><sup>3</sup>3http://www-supernova.lbl.gov. Because the charts were in postscript format, galactocentric separations had to be estimated visually, so each separation measurement has some degree of uncertainty associated with it, especially for SNe close to the cores of galaxies. Several of the SNe do have galactocentric separations reported in the circulars. Whenever possible, these numbers were used. The smallest group consists of eleven SNe that have been discovered in galaxies with Cepheid distances. These supernovae are important for the determination of the Hubble constant, as they set the absolute distance scale. The final group of SNe studied was the group of 27 local SNe used by Riess et al. 1998 to calibrate the MLCS and $`\mathrm{\Delta }m_{15}(B)`$ techniques. The Supernova Cosmology Project used only 18 local calibrators, nearly a subset of the Riess et al. sample, so are not studied separately here. Both sets are drawn from the Calán/Tololo SN survey (see Phillips et al. 1999), so they were discovered by the same methods, and should have nearly identical properties. This amounts to comparing the high redshift SNe (mostly) from one team with the low redshift SNe of another team, but this is not expected to be a problem. The two groups use similar discovery techniques for the high redshift SNe and nearly identical SNe for the low redshift calibrators, so the conclusions of this paper should be applicable to both groups. We have chosen the only data available at high redshift, and the most statistically significant sample of low redshift calibrators to study in this paper, and these happen to be from different research groups. ## 3 Results Histograms of the distribution of SNe with respect to PGD were generated for SNe discovered with CCDs locally, at high redshift, and SNe discovered photographically. These are presented in Fig. 1. There are not enough data points in the Riess et al. calibration sample to generate a meaningful histogram. SNe are presented with respect to raw PGD. Care must be taken in the interpretation of such data because it can be misleading and is impossible to properly normalize with the data available. As one travels farther out in PGD, the area swept out by each histogram bin increases. Corrections for this effect are not posible because there is no way to know the true distance of each SN from the center of its host galaxy (only the projected distance is known). Also, the inclination is not known for each galaxy, and the areal calibration would be different for face-on spirals than it would be for edge-on spirals. It is noteworthy that SNe Ia discovered on CCDs are found in greater numbers in bins covering the smallest area — those closer to the center of the galaxy. Additional unremovable forms of bias are present in the data. The projection of a three dimensional galaxy onto the two dimensional sky can cause interesting effects. High PGDs are more likely to be representative of true galactocentric distances. Consider an edge on spiral galaxy. SNe at high PGD are confined to a smaller range of distances perpendicular to the plane of the sky than SNe observed near the center of the galaxy. SNe with zero PGD may actually have a very high true galactocentric distance. The smaller size and faintness of galaxies at high redshift make them more susceptible to random errors in determining the PGD. These random errors can act like systematic errors because they tend to work in the same direction. SNe near the centers of galaxies may appear to have a larger PGD due to noise. It is far more unlikely for noise to lower a PGD. Human selection effects may also play a role in biasing the data. SNe near the centers of galaxies may be neglected at high redshift due to the difficulty of obtaining spectra of SNe contaminated with galaxy light. Each data set was sorted in order of increasing PGD and cumulative distributions were generated, given in Figure 2. The left axis can be interpreted as the fraction of SNe in each sample that are projected interior to a given galactocentric distance. The greatest vertical distance between two cumulative distribution curves is the Kolmogorov-Smirnov (K–S) statistic *D* (Press et al. 1988). This is used to determine the K–S probability that two samples were drawn from the same distribution. A K–S test was done on each of the four data sets with respect to each other and is presented in Table 1. The null hypothesis is that the two samples in each comparison were drawn from the same distribution. The numbers in Table 1 are the percent chance that the null hypothesis is true. A few special cases deserve comment. The probability that the sample of SNe discovered on photographic plates is drawn from the same distribution as the sample of SNe discovered on CCDs (at comparable redshifts, $`z<0.3`$) is $`4\times 10^5`$. As can be seen in Fig. 2, the difference is due to the lack of SNe discovered near the centers of galaxies on photographic plates, i.e. the Shaw Effect. In contrast, the K–S probability that local SNe discovered with CCDs and high–$`z`$ SNe discovered with CCDs are drawn from the same distribution is 58%. This number is probably artificially low due to the uncertainties in the positions of several local SNe very near the centers of galaxies as mentioned above. Indeed, when the sample is restricted to SNe with PGD of 3 kpc or greater, the corresponding K–S probability increases to 74%. Note that according to K–S statistics, these probabilities approaching unity do not mean the distributions are the same, but that they cannot be differentiated. Figures 1 and 2 illustrate another selection effect caused by the differences between CCD and photographic discovery of SNe. SNe in the outer regions of galaxies that are discovered by photographic means tend to be missed or omitted in CCD searches. This is usually assumed to be due to the fact that CCDs have a smaller field of view and may not sample the outer regions of galaxies. One would thus expect that high–$`z`$ SNe would not show this effect, because their host galaxies have a smaller apparent angular diameter. Figure 2 shows that the high–$`z`$ SNe presented in this study do show such a cutoff. This is probably due to the limited number of data points. Assuming SNe Ia arise from an exponential disk population with a scale length of 5.0 kpc, the probablility of discovering a SN farther than 21 kpc from the center of the galaxy is 1.5%. In a sample of 49 SNe, it is not surprising that no SNe are seen at a galactocentric distance greater than 21 kpc. In addition, at high redshift, SNe at large PGD may not be followed up (Clocchiatti, private communication). The sample of local SNe used to calibrate the high–$`z`$ SNe of Riess et al. (1998) was compared to the other groups of SNe. Twenty-four of twenty-seven of these local calibrators (89%) were discovered photographically, and are also members of the sample of photographically discovered SNe studied in this paper. Figure 2 shows that the cumulative distribution of the local calibrators used by Riess et al. matches one expected from a photographic sample. The K–S test gives the probability that the photographic sample and the Riess et al. calibrators are drawn from the same distribution as 94%. This number is given only for completeness, as the K–S test is not meaningful for significantly overlapping samples. The most striking result is that the K–S test gives a probability of only $`10^3`$ that the high–$`z`$ SNe are drawn from the same distribution as the local SNe used to calibrate them. The calibrators tend to be located farther from the centers of galaxies than the high–$`z`$ sample, so again the Shaw effect is largely to blame. The implications for the use of these SNe as standard candles are discussed below. Tables 2 and 3 reveal another subtle selection bias brought about by the Shaw Effect that has previously been overlooked. It is well known that SN discoveries in nearby galaxies are relatively unaffected by the Shaw Effect due to their large angular size. Of the eleven SNe with Cepheid distances, ten were located at a PGD of less than 10 kpc (see Table 2). This is particularly troubling given the findings of WHW and Riess et al. (1999) that SNe within this distance are on average 0.3 magnitudes brighter than supernovae located farther from the centers of their host galaxies. The SNe that tie the distant SNe to the Cepheid distance scale are drawn from a sample of potentially overluminous supernovae. This has implications for the determination of H<sub>0</sub> as discussed in the next section. In summary, the closest SNe (which are used to tie all other SNe to the Cepheid distance scale) are free from the Shaw Effect, but biased toward overluminous SNe. The intermediate SNe were largely discovered photographically, and are quite susceptible to the Shaw Effect. As a result, these supernovae (which are used as calibrators to train the MLCS technique and derive the light curve–decline relationship) are biased toward underluminous SNe. Finally, the high–$`z`$ SNe were discovered with CCDs and are relatively immune to the Shaw Effect. These results can easily be seen in Table 3. The median PGD for the photographic sample is 9.1 kpc, compared to a median of only 4.4 kpc for the CCD sample. The very local Cepheid calibrators have the smallest median PGD at only 3.2 kpc. ## 4 Conclusions Recently Hatano et al. (1998) proposed an alternative explanation for the Shaw Effect. They speculated that the Shaw Effect in Type II and Ibc SNe arises because extinction causes these supernovae to appear dim when projected onto the centers of galaxies, and dimmer SNe are harder to detect with increasing distance. Their model does not predict a paucity of SNe Ia near the centers of galaxies, and they were unconvinced that there is a real observational deficit of SNe Ia in this region. The data presented here confirm that the Shaw effect does indeed exist for photographically discovered Type Ia supernovae. These data support the traditional interpretation of the Shaw Effect — SNe are missed due to saturation of galaxy cores on photographic plates. These data provide an estimate for the magnitude of the Shaw Effect in SNe Ia. If we assume the CCD sample is free from bias (the smaller FOV bias is small as noted earlier), then 80% of SNe should be located within a PGD of 10 kpc (see Figure 2 and Table 3). In the photographic sample, 50 SNe have a PGD $``$ 10 kpc. If this represents 20% of the sample, then the total sample should be $``$ 250 SNe; however, there are only 105 SNe in the sample, so $``$ 145 (58%) are missing. Thus nearly 60% of SNe Ia are lost near the centers of galaxies on photographic plates due to the Shaw Effect. This is striking because it is higher than most estimates of the Shaw Effect for all SNe, despite the fact that SNe Ia are $``$ 1.5 times brighter on average than SNe II. This is the key piece of evidence that saturation (not extinction) is to blame for the paucity of SNe. It would be very interesting to do a relative comparison of the samples of Type II and Type b/c SNe discovered with CCD’s compared to Type Ia SNe to explore the relative concentrations toward the centers of galaxies. The high redshift SN sample appears to be relatively free from selection bias in terms of separation from the center of the galaxy. These SNe were discovered with CCDs and show a distribution with respect to galactocentric distance similar to that of local SNe discovered with CCDs. We find no evidence that high–$`z`$ SNe are selectively discovered farther from the centers of galaxies. The high–$`z`$ SN studies may be sensitive to selection effects. According to Riess et al. (1998), “we must continue to be wary of subtle selection effects that might bias the comparison of SNe Ia near and far.” They also add, “It is unclear whether a photographic search selects SNe Ia with different parameters or environments than a CCD search or how this could affect a comparison of samples. Future work on quantifying the selection criteria of the samples is needed.” This raises a potential area of concern when using SNe Ia as distance indicators. The light curve decline – absolute magnitude relationship used to calibrate the brightness of high–$`z`$ SNe is derived mainly from local SNe discovered photographically, as is the case in Riess et al. (1998) and Perlmutter et al. (1999). These SNe tend to be located farther from the centers of galaxies than the SNe in the high–$`z`$ sample. The calibrators are thus drawn from a sample subject to selection bias that is both more uniform in luminosity and dimmer than average. Because the calibrators are likely more homogeneous as a group than the high–$`z`$ SNe, they may not be able to effectively correct for the light curve decline relationship over all of parameter space. This work has shown that there are issues of selection bias to be considered, but how much these issues affect the use of SNe Ia as distance indicators remains unclear. Fig. 6c of Riess et al. (1999) shows that after SN magnitudes are corrected for light curve shape using the MLCS technique there is no apparent trend of absolute V magnitude at maximumn with PGD. In this case the calibration process seems to work adequately. No data is available to indicate how well other calibration methods, particularly $`\mathrm{\Delta }m_{15}(B)`$, (used by Riess et al. 1998 in addition to MLCS) and the stretch factor method (used by Perlmutter et al. 1999), can compensate for changes in the absolute magnitudes of SNe with PGD. Since these calibration methods all rely on the same principle — using the shape of the light curve to correct for variations in luminosity — one might expect that all will work similarly. It is also not known how the calibration process affects the trend reported by WHW of increased scatter in the colors of SNe at maximum light with decreasing PGD. Future work is needed to assure that all methods of calibration of SNe Ia can remove trends with PGD in all bandpasses. No previous studies have found significant problems with the calibration process, so the results of Riess et al. (1998), and Perlmutter et al. (1999) do not appear to be at risk due to selection bias. Evidence for a local void (Zehavi et al. 1998), already marginal, may be more sensitive to subtle effects. The PGD-magnitude effect when combined with the Shaw effect may explain several unusual findings reported recently. Tammann et al. (1996) and Tammann (1998) noted that after SNe luminosities are corrected via a light curve decline relationship, local SNe with Cepheid distances are brighter than average, and this is the opposite than one would expect from a magnitude limited sample (the farthest SNe should be brightest because we could not see the faint ones). This discrepency has been used to argue that there is a problem with light curve decline relationships. From the result presented here – that ten of eleven SNe with Cepheid distances have PGD $`<10`$ kpc – these SNe can be expected to be overluminous. Two of the SNe with reported Cepheid distances, 1980N and 1992A, do not actually have a Cepheid distance determined to their host galaxies, but are members of the Fornax cluster, for which a Cepheid distance of $`(mM)_0=31.35`$ is available by assuming that the galaxy NGC 1365 is a member of the cluster (Madore et al. 1999). These two SNe appear to be too dim, by 0.3 mag, compared to the Calán/Tololo sample of SNe Ia, leading Suntzeff et al. (1999) to speculate that NGC 1365 may be foreground to Fornax by 0.3 mag. Interestingly, these SNe have the highest PGD of the SNe in Cepheid galaxies, at 20.0 and 9.3 kpc, respectively, both outside the distance WHW identify as the cutoff beyond which SNe Ia are seen to be 0.3 mag dimmer on average. A possible explanation for the faintness of SN 1980N and SN 1992A is their location in their host galaxies, rather than an error in the distance to Fornax. It should also be noted, however, that Saha et al. (1999) derive a distance modulus to NGC 1316 of $`(mM)_0=31.84`$, by assuming that SN 1980N in NGC 1316 is similar to SN 1989B, which would make SN 1980N too bright. The fact that the Cepheid calibrators are drawn from a potentially overluminous sample of SNe is troubling. This could affect determinations of the Hubble constant that rely on using these very local SNe for calibration purposes. The fact that different biases operate on different distance scales calls into question the variation of H<sub>0</sub> with distance reported by Tammann (1998), as no second parameter correction was used. More data is needed on the galactic radial distribution of high and low redshift SNe. Given the discovery rate of high–$`z`$ SNe and the fact that they presumably do not face the discovery bias related to the small size of a CCD relative to the host galaxy, the statistics of the intrinsic distribution of high–$`z`$ SNe may soon be known more precisely than those of the low redshift sample. In an ideal world, the luminosity of high–$`z`$ SNe would only be calibrated with SNe discovered using CCDs. That is not possible at the present time, so early work indicating that light curve decline – absolute magnitude relationships can effectively remove the effects of selection bias should be continued and strengthened. ###### Acknowledgements. The authors thank the anonymous referee for his or her helpful comments and Adam Riess and Alan Sandage for comments and perspective. This research was supported in part by NSF Grant 95-28110, a grant from the Texas Advanced Research Program, and by NASA through grant HF-01085.01-96A from the Space Telescope Science Institute which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555.
no-problem/9908/nucl-th9908065.html
ar5iv
text
# 1 Introduction ## 1 Introduction Judah M. Eisenberg was elegant in appearance as well as in expression. His taste for elegance carried over into his research, and hence it was most natural that his attention should also be drawn to Skyrme’s topological soliton model for the baryons. He published 21 papers on the application of this model, the mathematical beauty of which trancends its phenomenological utility. In the non-perturbative region, which comprises most low energy baryon structure and all of nuclear structure, the large color limit of QCD has proven of great utility . In this limit those Feynman diagrams, which have the largest $`N_c`$ factor in dominate the $`S`$-matrix. Those diagrams involve only planar gluons, which then may be replaced by $`q\overline{q}`$ pair lines, and as a consequence all surviving diagrams admit a meson exchange interpretation. In the large $`N_c`$ limit counting of the $`N_c`$-factors reveals mesons to be stable and non-interacting. Baryons are different. They are color singlets of $`N_c`$ quarks, and accordingly their mass scales as $`N_c\mathrm{\Lambda }_{QCD}`$, where $`\mathrm{\Lambda }_{QCD}`$ is the (only) dimensional QCD scale factor. Their radius remains proportional to $`1/\mathrm{\Lambda }_{QCD}`$ and $`N_c`$ independent. Accordingly the quark density with the baryon grows beyond bound, and a Hartree approximation becomes appropriate . Finally as the meson-baryon coupling constants in general are proportional to $`\sqrt{N_c}`$, mesons are strongly coupled to baryons in the large $`N_c`$ limit. There are two approaches to describe the baryons in a large $`N_c`$ limit. The first is to employ a constituent quark model description based on appropriately symmetrized products of $`N_c`$ quark wave functions in the Hartree approximation. The other is to construct the baryons as topologically stable soliton solutions to a chiral Lagrangian of meson fields, with the general form $$_{meson}=N_c_p(\frac{\varphi }{\sqrt{N_c}}),$$ $`(1)`$ where $`_p`$ is polynomial of meson fields and their gradients, which satisfies the (nontrivial) stability requirements. The Skyrme model , and its generalizations are generic examples of the latter approach. ## 2 Large $`N_c`$ Operator Algebra Among mesons the pions stand out because their role as the Goldstone bosons of the spontaneously broken approximate chiral symmetry of QCD. As such their coupling to hadrons has to vanish with 4-momentum. The pion coupling to fermions is accordingly $$=\frac{1}{f_\pi }A^{\mu a}_\mu \pi ^a,$$ $`(2)`$ where $`f_\pi `$ is the pion decay constant $`(\sqrt{N_c})`$, and $`A^{\mu a}`$ is the axial vector of the fermion, which in the case of a baryon scales as $`N_c`$. In the large $`N_c`$ limit, the baryons propagator reduces to the static propagator, and $`\pi `$-baryon scattering becomes recoilless. Consequently the $`\pi `$-baryon scattering amplitude only involves bilinear combinations of the space components of the axial current, from which it proves convenient to separate a factor $`N_c`$ as $$A^{ia}=gN_cX^{ia},$$ $`(3)`$ where $`g`$ is a coupling constant introduced for convenience. For the operators $`X^{ia}`$ the following $`N_c`$ expansion proves useful : $$X^{ia}=X_0^{ia}+\frac{1}{N_c}X_1^{ia}+\frac{1}{N_c^2}X_2^{ia}+\mathrm{}$$ $`(4)`$ The lowest order operator $`X_0^{ia}`$ and the spin- and isospin operators then satisfy a contracted SU(4) algebra. This is the algebra of the Skyrme model, which is made explicit by the identification $$X_0^{ia}=\frac{1}{2}Tr\{A\sigma ^iA^{}\tau ^a\},$$ $`(5)`$ A being the SU(2) rotational collective coordinate used in the spin-isospin quantization of the Skyrme Hamiltonian . Baryon operators may then described systematically in terms of the operators $`X_n^{ia}`$ or alternatively in terms of the SU(4) generators spin, isospin and spin-isospin $`G^{ia}`$, where the relation to the $`X_n^{ia}`$:s is given by the limiting relation $$\underset{N_c\mathrm{}}{lim}\frac{G^{ia}}{N_c}=X_0^{ia}.$$ $`(6)`$ The connection to the constituent quark model obtains by expression of the SU(4) generators in terms of corresponding quark operators: $$J^i=\underset{l=1}{\overset{N_c}{}}q_l^{}\frac{\sigma ^i}{2}q_l,I^a=\underset{l=1}{\overset{N_c}{}}q_l^{}\frac{\tau ^a}{2}q_l,$$ $$G^{ia}=\underset{l=1}{\overset{N_c}{}}q_l^{}\frac{\sigma ^i}{2}\frac{\tau ^a}{2}q_l.$$ $`(7)`$ Thus in the large $`N_c`$ limit the constituent quark model and the Skyrme model give equivalent results. ## 3 Phenomenological considerations The formalism outlined above may be employed to derived systematic $`1/N_c`$ expansions for baryon operators as masses, magnetic moments . Relations between the available two-body operators allows expression of the baryon mass operator as $$M=m_0N_c+m_2\frac{1}{N_c^2}\stackrel{}{J}^2+m_3\frac{1}{N_c^3}\stackrel{}{J}^4+..,$$ $`(8)`$ where $`\stackrel{}{J}`$ is the spin-operator. To order $`1/N_c^2`$ this agrees with the Skyrme model result $$M=M_s+\frac{1}{2\mathrm{\Omega }_s}\stackrel{}{J}^2,$$ $`(9)`$ where $`M_s`$ and $`\mathrm{\Omega }_s`$ are the mass and moment of inertia of the soliton. Returning to the quark model representation, (7) it has been found that among the two-body operator combinations of $`J^i`$, $`I^a`$ and $`G^{ia}`$, the only significant combination for the mass operator is the operator $$\underset{i<j}{}\stackrel{}{\sigma }_i\stackrel{}{\sigma }_j\stackrel{}{\tau }_i\stackrel{}{\tau }_j,$$ $`(10)`$ and its generalization SU(3) flavor . This is consistent as it is this operator and the scalar unit operator, which have the highest $`N_c`$-scaling factors . As this operator appears in the pion exchange and multipion exchange interaction between constituent quarks, it suggests that in the region between the chiral restoration scale $`\mathrm{\Lambda }_\chi ^11/4f_\pi `$ and the confinement scale $`1/\mathrm{\Lambda }_{QCD}`$ the effective dynamical description of baryons is in terms of constituent quarks that interact by exchanging pions . The reason for the effectiveness of the operator (8) in organizing the baryon spectrum in agreement with the empirical one is that it is the only operator combination that is able to achieve the correct ordering of the positive and negative parity states in the spectrum. Any monotonic confining interaction would order the states in baryon spectrum in shells of alternating parity. Yet the lowest excited states in the nucleon spectrum are successively the $`N(1440),\frac{1}{2}^+`$ and the $`N(1535),\frac{1}{2}^{}`$. As these states (and the nucleon) all have mixed color-spin symmetry $`[21]_{CS}`$, the color-spin dependent perturbative gluon exchange model for the hyperfine interaction cannot reverse their normal ordering and bring them into agreement with the empirical ordering. On the other hand the $`N(1535),\frac{1}{2}^{}`$ state has mixed flavor-spin symmetry $`[21]_{FS}`$, whereas the nucleon and the $`N(1440),\frac{1}{2}^+`$ state have complete flavor-spin symmetry $`[3]_{FS}`$. Therefore, in combination with a sufficiently strong orbital matrix element, the operator (8) is able to reverse the normal ordering, and bring the spectrum into agreement with experiment . This argument carries over to all flavor sectors of the baryon spectrum, with exception of the $`\mathrm{\Lambda }`$-spectrum, where the negative flavor singlet $`\mathrm{\Lambda }1405\mathrm{\Lambda }1520`$ is the lowest multiplet, which however, also only can be understood with the operator form (8). The interaction (8) may be interpreted as being due to pion and multipion exchange between the constituent quarks. Two-pion exchange is required to cancel out the tensor interaction of one-pion exchange, as the tensor interaction would otherwise imply substantial spin-orbit splittings in the $`P`$-shell, in conflict with experiment . With suitable parameter choices the inclusion of two-pion exchange also allows a cancellation of the spin-orbit interaction that arises with a linear scalar confining interaction. The constituent quark model, with quarks interacting by pion exchange, is usually referred to as the chiral quark model. The dynamical interpretation resembles that of the Skyrme model, Hence, and in view of the discussion in section 2 above, it is no surprise that the reversal of the normal ordering of the states in the baryon spectrum also obtains in the Skyrme model. This model has long been known to imply very low lying vibrational states . ## 4 Heavy Flavor Baryons and Pentaquarks The quark model describes baryons as 3-quark systems. The Skyrme model describes baryons as formed of $`(q\overline{q})^n`$ \- i.e. mesons - in the field of a soliton, which carries the baryon number. Hyperons are then best described as bound states of soliton and $`K`$, $`D`$ and $`B`$ mesons respectively . This provides a unified description of normal hyperons and ”pentaquarks”. While a hyperon is described as a bound state of a soliton and a heavy flavor meson, the corresponding pentaquark is described as a bound state of soliton and the corersponding heavy flavor antimeson . Note that the nomenclature is due to the non-symmetric description by the constituent quark model of the former as $`qqQ`$ states and the latter as $`qqqq\overline{Q}`$ states, where $`q`$ and $`Q`$ denote light and heavy flavor quarks respectively. Both the chiral quark model, extended to broken $`SU(N_F)`$ , and the bound state soliton model describe the ground state heavy flavor hyperons well . But only the latter yields absolute predictions for the pentaquark energies without further parameters or assumptions, as the difference between hyperon and pentaquark energies arises solely from the sign of the Wess-Zumino term in the effective meson-soliton interaction . The existence or non-existence of pentaquarks is interesting, because it depends very much on the form of the hyperfine interaction between quarks. If that interaction has the form of perturbative gluon exchange, the lowest energy pentaquark contains a strange quark, and has negative parity . The bound state soliton model and the chiral quark model in contrast predict that the lowest energy - and definitely stable - pentaquark is non-strange and has positive parity. Only the former has been searched for experimentally, although with somewhat inconclusive results . Experimental identification of the latter would be very informative, and would automatically explain the non-existence of the $`H`$-particle, which hasn’t been found in spite extensive effort . ## 5 Nuclei Most of Judah Eisenberg’s work on applications of the Skyrme model dealt with nuclei. This work employed Skyrme’s product ansatz for solitons with baryon number larger that 1 : $$U(r;\stackrel{}{r}_1,\mathrm{}\stackrel{}{r}_A)=\mathrm{\Pi }_{i=1}^AU(\stackrel{}{r}\stackrel{}{r}_1)$$ $`(9)`$ When this ansatz is inserted into the Lagrangian density of the Skyrme model, the Lagrangian separates into a sum of single nucleon terms and interaction terms involving $`2,\mathrm{\hspace{0.17em}3},\mathrm{}A`$ nucleons. When integrated over $`\stackrel{}{r}`$ the latter yield models for the $`2`$-, $`3`$-, …$`A`$-nucleon interactions. For large internucleon separations these terms have the same form as the corresponding pion-exchange interactions that obtain with more conventional chiral Lagrangians . This is illustrated by the derivation of the three-nucleon interaction based on the product ansatz by Eisenberg and Kälbermann . For large interparticle separations this interaction reduces to the conventional two-pion exchange three-nucleon interaction, that involves an intermediate $`\mathrm{\Delta }_{33}`$ resonance in the sharp resonance approximation . The appearance of the sharp resonance propagator $`i(m_\mathrm{\Delta }m_N)`$, in this expression is a direct consequence of the large $`N_c`$ mass formula (9). The product ansatz (9) does not, however, provide a good approximation to the minimum energy configuration of 8 skyrmions with $`B>1`$. The minimum energy solutions have an interesting topology, which hitherto has not yielded to analytical treatment . This situation has now changed by the discovery of elegant rational map approximations to the minimal energy solutions . The rational map approximation for the $`B=n`$ skyrmion takes the form $$U(\stackrel{}{r})=e^{i\stackrel{}{\tau }\stackrel{}{\pi }_nF(r)},$$ $`(10)`$ where the ”chiral angle” $`F(r)`$ depends only on the distance to the center, and $`\stackrel{}{\pi }_n`$ is defined as $$\stackrel{}{\pi }_n=\frac{1}{1+|R_n(z)|^2}\{2Re[R_n(z)],2Im[R_n(z)],1|R_n(z)|^2\}.$$ $`(11)`$ Here $`R_n(z)`$ is the rational map for baryon number $`n`$, and $`z`$ is defined as $`z=tanv/2e^{i\rho }`$. For $`n=1`$, $`R(z)=z`$, and (10) reduces to Skyrme’s hedgehog solution. For larger $`n`$ $`R(z)`$ are simple rational functions of $`z`$, the simplest case being $`n=2`$ for which $`R_2(z)=z^2`$. The rational maps open the door to more realistic applications to nuclei based on the Skyrme model, than what hitherto has been possible. As an example the question of the existence of bound states between $`\eta `$-mesons and nuclei may be addressed and shown to be likely . What is still wanting, however, are functional forms that connect the rational maps approximations (10) to the product ansatz (9), which is appropriate for large separations. ## 6 Judah Juhah would have enjoyed the elegance of the rational maps. Of his style in expression, here he is in a letter, dated November 12, 1973, in Charlottesville: “Stanley Hanna and I were very much hoping that you might be able so see your way clear to give a talk at the \[1974 photonuclear Gordon\] conference. In particular, we wondered if you would be willing to review … . Of course, it would be appropriate to slant your discussion towards aspects of this problem, which would be of particular relevance for the community of photonuclear physicists”. Obviously I accepted. That photonuclear conference, which Judah chaired, will be long remembered by the participants, as one of the (unscheduled) evening talks was the televised resignation speech of R. M. Nixon.
no-problem/9908/cond-mat9908280.html
ar5iv
text
# Linear Field Dependence of the Normal-State In-Plane Magnetoresistance of "Sr"₂⁢"RuO"₄ \[ ## Abstract The transverse and longitudinal in-plane magnetoresistances in the normal state of superconducting $`\text{Sr}_2\text{RuO}_4`$ single crystals have been measured. At low temperatures, both of them were found to be positive with a linear magnetic-field dependence above a threshold field, a result not expected from electronic band theory. We argue that such behavior is a manifestation of a novel coherent state characterized by a spin pseudo gap in the quasi-particle excitation spectrum in $`\text{Sr}_2\text{RuO}_4`$. \] Since the discovery of superconductivity in $`\text{Sr}_2\text{RuO}_4`$, isostructural with high-$`T_\text{c}`$ cuprate $`\text{(La,Sr)}_2\text{CuO}_4`$, this material has emerged as a new focus of superconductivity research. There has been a growing body of experimental evidence for a spin triplet (odd-parity) pairing in $`\text{Sr}_2\text{RuO}_4`$, including results obtained in muon spin relaxation, NMR Knight shift, electrical transport, proximity effect, and specific heat measurements. In the normal state, a highly anisotropic Fermi liquid behavior, characterized by a $`T^2`$-dependence in the in- and out-of-plane resistivity ($`\rho _{ab}`$ and $`\rho _c`$, respectively) and a cylindrical Fermi surface, was found for $`\text{Sr}_2\text{RuO}_4`$ at low temperatures. The enhancements of electronic specific heat and spin susceptibility are very similar to those found in $`{}_{}{}^{3}\text{He}`$, suggesting that $`\text{Sr}_2\text{RuO}_4`$ is a strongly correlated system. At higher temperatures, however, the situation is not very clear. While $`\rho _c`$ undergoes a nonmetallic (d$`\rho _c/dT<0`$) to metallic (d$`\rho _c/dT>0`$) crossover around 130K, a similar feature is not seen in $`\rho _{ab}`$, a phenomenon yet to be fully understood. Chemically, $`\text{Sr}_2\text{RuO}_4`$ is positioned in the vicinity of both ferromagnetic and antiferromagnetic orderings. In the Ruddlesden-Popper homologous series $`\text{Sr}_{n+1}\text{Ru}_n\text{O}_{3n+1}`$, as the single-layer (n $`=1`$) member, $`\text{Sr}_2\text{RuO}_4`$ is a paramagnetic metal with no apparent local moment. The infinite-layer ($`\text{n}=\mathrm{}`$) member, $`\text{SrRuO}_3`$, is a ferromagnetic metal. The n = 2 member of the series, $`\text{Sr}_3\text{Ru}_2\text{O}_7`$, which has been reported to be ferromagnetic by one group and antiferromagnetic by others, should be magnetic as well. On the other hand, if $`\text{Sr}^{2+}`$ ions in $`\text{Sr}_2\text{RuO}_4`$ are replaced by $`\text{Ca}^{2+}`$, the resultant compound $`\text{Ca}_2\text{RuO}_4`$, which is isostructural and isoelectronic to $`\text{Sr}_2\text{RuO}_4`$, is an antiferromagnetic insulator. Being in the vicinity of these very different behaviours, $`\text{Sr}_2\text{RuO}_4`$ is likely to exhibit unusual electronic and magnetic properties. The normal-state properties of $`\text{Sr}_2\text{RuO}_4`$ can be probed by magnetoresistance (MR) measurements. The normal-state MR of $`\text{Sr}_2\text{RuO}_4`$ was first measured at 20mK in the context of determining the Fermi surface of this material through the Shubnikhov-de Haas oscillations. However, no field dependence of MR was identified in that work. More recently, Hussey et al. have measured both c-axis and in-plane normal-state MR of $`\text{Sr}_2\text{RuO}_4`$, focusing primarily on the c-axis MR ($`\mathrm{\Delta }\rho _c/\rho _c=(\rho _c(H)\rho _c(0))/\rho _c(0)`$) at relatively high temperatures. The in-plane MR ($`\mathrm{\Delta }\rho _{ab}/\rho _{ab}=(\rho _{ab}(H)\rho _{ab}(0))/\rho _{ab}(0)`$) was measured only in transverse configuration for one sample, without a detailed analysis of its field dependence over the entire temperature range. Longitudinal in-plane MR results were not available in literature prior to the present study. In this paper, we report results of our systematic study of the normal-state in-plane MR for $`\text{Sr}_2\text{RuO}_4`$ in both transverse and longitudinal configurations. Our work has revealed previously unknown features in in-plane MR, strongly suggesting a novel behavior in the normal state of $`\text{Sr}_2\text{RuO}_4`$. Single crystals used in this study were grown by a floating-zone method with details described elsewhere. Resistance measurements show a superconducting transition temperature ($`T_c`$) of 0.84K. MR measurements were carried out in a $`{}_{}{}^{3}\text{He}`$ cryostat. The temperature was measured using a Lakeshore Cernox 1030 thermometer with relative temperature corrections, due to magnetic field, typically 0.15% at 4.2 K and -0.023% at 77.8K and 8.0T. For in-plane MR and Hall measurements, we used four rectangular shaped single crystals with dimensions around $`1.2\times 1.0\times 0.01`$, $`0.9\times 0.4\times 0.08`$, $`0.8\times 0.2\times 0.07`$, and $`0.6\times 0.3\times 0.05\text{mm}^3`$, respectively. For each sample, two current contacts covering the opposite ends and four voltage contacts on the two sides of the crystal were prepared. All $`\text{RuO}_2`$ layers were electrically shorted along the c-axis to ensure a homogeneous current distribution. For transverse and longitudinal MR measurements, the magnetic field $`H`$ was applied perpendicular and parallel to the injected current $`I`$, respectively. In order to exclude the Hall contribution to MR, only the symmetric part of $`\mathrm{\Delta }\rho _{ab}(H)=\rho _{ab}(H)\rho _{ab}(0)`$ under field reversal was included. For in-plane Hall measurements, the magnetic field was applied parallel to $`c`$-axis with a current bias applied along the $`ab`$-plane. The Hall voltage $`V_\text{H}`$, which contains only the asymmetric contributions under the field reversal, was found to vary linearly with $`H`$ up to 5T. By fitting $`V_\text{H}(H)`$ data in the linear regime using $`V_\text{H}=R_\text{H}HI/d`$ ($`d`$ is the thickness of the sample), the in-plane Hall coefficient $`R_\text{H}`$ was obtained. The transverse in-plane MR, $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ ($`HI`$), between 0 and 7.3T at various temperatures is shown in Fig. 1. Similar results have been obtained in separate samples. $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ is seen to be positive, growing rapidly in magnitude with decreasing $`T`$. At low fields, $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ at a fixed temperature is very small, which can be described by the sum of a linear and a quadratic term. When $`H`$ exceeds a threshold value $`H_0`$, $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}=K_\text{S}(HH_0)`$, where $`K_\text{S}`$ and $`H_0`$ are functions of temperature. The solid lines in Fig. 1 represent fits to the data above $`H_0`$ using this form. The temperature dependence of $`H_0`$ is shown in the inset of Fig. 1. We note that in Ref. 20, results on $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$, plotted against $`H^2`$, were presented. While $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ might be quadratic in $`H`$ at the two highest temperatures shown $`(`$56K and 82K$`)`$, a deviation from this behavior can be seen at all other temperatures. (We have re-plotted the low temperature data against $`H`$ and found both the magnitude and the field dependence of $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ given in Ref. 20 are in agreement with those obtained in the present study at overlapping temperatures.) Within the band theory for solids, MR is proportional to $`(H/\rho (0))^2`$ in the low-field limit. Linear MR can be expected only in some special circumstances. For example, a square Fermi surface can lead to linear MR due to the presence of the sharp corners. Local-density approximation calculations have shown that the Fermi surface of $`\text{Sr}_2\text{RuO}_4`$ consists of three roughly cylindrical-like sheets with no such sharp corners, thus excluding this from being the origin of the observed linear MR. For systems with multi-band electronic structure involving two types of charge carriers, the field dependence of MR can be written as $`\mathrm{\Delta }\rho /\rho (0)=\text{a}H^2/(\text{b}+\text{c}H^2)`$ where a, b, and c are positive, field-independent quantities determined by the relaxation rates of each type of carriers. It is clear that this expression will not lead to $`\mathrm{\Delta }\rho /\rho (0)H`$ as observed experimentally. In some elemental metals (such as K, In, and Al), linear MR has been found and attributed to the boundary and disorder effects. Since superconductivity in $`\text{Sr}_2\text{RuO}_4`$ is extremely sensitive to the amount of disorder, the linear $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$, found reproducibly in superconducting $`\text{Sr}_2\text{RuO}_4`$ with $`\rho _{ab}(0)1\mu \mathrm{\Omega }`$cm, is unlikely due to effects of the impurities and/or structural defects. Linear MR was observed previously in heavy fermion $`\text{CeCu}_6`$ at very low temperatures. In another heavy fermion compound, $`\text{UPt}_3`$, the MR was found to be positive and proportional to $`H^{1.25}`$. The observed linear MR was interpreted as being a consequence of the opening of a pseudo gap in a coherent state formed at low temperatures. In this picture, the pseudo gap is suppressed by the magnetic field, resulting in enhanced scattering rates and therefore positive MR. NMR studies of $`\text{Sr}_2\text{RuO}_4`$ provided the first hint that a coherent normal state may also be present in this material at low temperatures . The emergence of such a state is signaled by a sharp change of slope in $`1/T_1`$ vs. temperature curve for both $`{}_{}{}^{101}\text{Ru}`$ and planar $`{}_{}{}^{17}\text{O}`$ sites around 80K. Our in-plane MR results also indicate that this temperature is special for $`\text{Sr}_2\text{RuO}_4`$. As shown in Fig. 2, a sign change from positive to negative in $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ is seen between 70-80 K, similar to what has been observed in c-axis MR. It should be noted that the range of the data scattering in $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ at 7.3 T is around 0.004%. The corrections in temperature due to the MR of the thermometer is around -0.016% at 87 K and 8.0 T, which cannot lead to the observed sign change in $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$. While the physical origin of the sign reversal in MR has not been understood at the present time, it is not inconceivable that this is associated with the emergence of a coherent normal state in $`\text{Sr}_2\text{RuO}_4`$ at low temperatures. As will be argued later, such a state may be accompanied by the opening of a pseudo gap in the quasi-particle excitation spectrum. Similar to $`\text{CeCu}_6`$, an applied magnetic field suppresses the pseudo gap, leading to linear in-plane MR for $`\text{Sr}_2\text{RuO}_4`$. The in-plane Hall coefficient $`R_\text{H}`$ of $`\text{Sr}_2\text{RuO}_4`$, should be subject to the same scattering processes as that for the in-plane MR. An interesting question is whether $`R_\text{H}`$ will reflect the presence of a pseudo gap in the quasi-particle excitation spectrum. The in-plane $`R_\text{H}`$ was measured for one of our samples in the same magnetic field and current configuration as that for the transverse in-plane MR. The results are shown in Fig. 3. The Hall coefficient is negative at high temperatures, changing its sign around 130K, which is roughly where $`\rho _c`$ undergoes a nonmetallic-metallic crossover. It reaches a maximum between 70-80K before changing its sign again around 30K. Prior to the present work, the Hall coefficient $`R_\text{H}`$ of $`\text{Sr}_2\text{RuO}_4`$ has been measured on crystals prepared by different groups. Our results agree well with those published earlier, indicating that features shown in $`R_\text{H}(T)`$ are intrinsic. In Ref. 19, an isotropic-$`\mathrm{}`$ approximation within the multi-band picture of $`\text{Sr}_2\text{RuO}_4`$ was used to calculate the low temperature limit ($`T<1`$K) of $`R_\text{H}`$ with band parameters obtained from quantum oscillations. The calculated result agrees with the experimental value. On the other hand, a manybody-theory-based calculation of $`R_\text{H}(T)`$ has also been carried out, yielding a result that is consistent with experimental observations. This suggests that an account for behaviors in $`R_\text{H}`$ beyond band theory is still possible, especially for $`T>1`$K where the isotropic-$`\mathrm{}`$ approximation did not seem to work. It is interesting to note that the characteristic temperature where $`R_\text{H}`$ shows a maximum coincides with that below which we believe the pseudo gap emerges. Whether the maximum in $`R_\text{H}`$ is another signature for the opening of a pseudo gap in $`\text{Sr}_2\text{RuO}_4`$, and how strongly the presence of such a gap will affect $`R_\text{H}`$, are issues to be resolved. In order to infer the relative contribution of the orbital and spin motions to MR and the physical origin of this pseudo gap, we have measured the longitudinal in-plane MR of $`\text{Sr}_2\text{RuO}_4`$. As shown in Fig. 4, the longitudinal in-plane MR $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ also revealed a linear behavior above $`H_0`$, which increases with increasing temperature (see the inset of Fig. 4). More striking is that the magnitude of $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ is greater than that of $`\mathrm{\Delta }\rho _{ab}^{}/\rho _{ab}`$ below approximately 10K, indicating that the spin contribution to the in-plane MR is important. Therefore, the pseudo gap may be of spin origin. Results obtained in NMR <sup>17</sup>O Knight shift measurements in $`\text{Sr}_2\text{RuO}_4`$ point to the same direction. When the <sup>17</sup>O Knight shift results were decomposed into spin susceptibilities of three Ru d-orbitals, $`d_{xy}`$, $`d_{xz}`$, and $`d_{yz}`$, a broad maximum around 40K was found. This is consistent with uniform susceptibility results where a broad peak was seen around the same temperature. The decrease of spin susceptibility at low temperatures may then be taken as an indication of the opening of a pseudo gap in the spin excitation spectrum. It has been suggested theoretically that the spin fluctuations in $`\text{Sr}_2\text{RuO}_4`$ are predominantly ferromagnetic. Experimentally, $`1/T_1`$ obtained from the NMR measurements for both O(1) and Ru sites shows an identical temperature dependence. In general, $`1/T_1`$ can be expressed by the sum of q-dependent imaginary part of the dynamical electron susceptibility with appropriate form factors, which should have different q-dependence for <sup>101</sup>Ru and planar <sup>17</sup>O sites, leading to different temperature dependence in their respective $`1/T_1`$. However, if the dynamical electron susceptibility has a sharp peak around $`𝐪=(0,0)`$, corresponding to ferromagnetic fluctuations, then the above mentioned behavior in $`1/T_1`$ can be explained. On the other hand, this scenario will lead to negative MR, which contradicts the experimental observation of positive MR. Results obtained in inelastic neutron scattering (INS) measurements carried out recently on $`\text{Sr}_2\text{RuO}_4`$ have resolved this apparent contradiction between NMR and magnetotransport results. In this experiment, incommensurate magnetic spin fluctuations located at $`𝐪_0=(\pm 0.6/a,\pm 0.6/a,0)`$, originally predicted by band calculations, were found at low temperatures. The intrinsic q-width for these incommensurate fluctuations was found to be narrow ($`\mathrm{\Delta }𝐪=0.13\pm 0.02`$Å at 10K). The temperature dependence of $`1/T_1`$ observed in NMR can be reconstructed from the INS results, confirming that the dominating spin fluctuations are incommensurate rather than ferromagnetic as suggested in Ref. 26. The energy dependence of the imaginary part of the dynamic susceptibility at $`𝐪_0`$, which measures the dissipation of the spin fluctuations, shows a gradual drop below 7meV, indicating the presence of a gap in the spin excitation spectrum. We note here that this energy scale (7 meV) agrees well with the value of the pseudo gap inferred from the temperature dependence of other physical quantities such as $`1/T_1`$ and MR. Physical insight into the nature of the normal state of $`\text{Sr}_2\text{RuO}_4`$ can be obtained from the high-pressure experiments. It was shown that the superconducting transition temperature $`T_c`$ was suppressed by an applied pressure $`p`$. By extrapolating results obtained up to 1.2 GPa, $`T_c`$ is expected to become zero around $`p_c`$ 3 GPa. In normal state, $`\rho _{ab}`$ was found to vary non-monotonically with $`p`$ at low $`T`$. With increasing $`p`$, $`\rho _{ab}`$ first increases, and then decreases as $`p`$ was increased further above a threshold value, close to 3 GPa. At 8 GPa, $`\rho _{ab}`$ was found to show a $`T^{4/3}`$-dependence, characteristic of a two-dimensional electronic system close to a ferromagnetic quantum critical point. Based on this, we propose the following phase diagram for $`\text{Sr}_2\text{RuO}_4`$, shown schematically in Fig.5. At $`T`$ = 0 K, a direct superconductor (SC)-to-ferromagnetic metal (FM) transition, tuned by pressure, occurs at $`p=p_c`$ 3 GPa. At low $`T`$ and low $`p`$, $`\text{Sr}_2\text{RuO}_4`$ is superconducting. A pseudo gap regime emerges above $`T_c`$, similar to the quantum spin disordered regime discussed in quantum critical phenomena. As $`p`$ is further increased, the pseudo gap is suppressed, leading to an increase in $`\rho _{ab}`$, as observed experimentally. When the system enters the fluctuating regime at a finite $`T`$, $`\rho _{ab}`$ decreases with the increasing $`T`$ since the system is driven towards a ferromagnetic state. We suggest that the magnetic field similarly suppresses the pseudo gap, resulting in a positive MR. In this picture, as $`T`$ increases, a larger threshold field ($`H_0`$) is needed to overcome the thermal smearing so that linear MR can be observed, consistent with our experimental observations. A pseudo gap has been observed in high-$`T_c`$ cuprate superconductors. However, the origin and physical consequences of pseudo gap in high-$`T_c`$ materials may be different from that in $`\text{Sr}_2\text{RuO}_4`$. The in-plane MR observed in $`\text{(La,Sr)}_2\text{RuO}_4`$ is positive, but not linear in field dependence. When the pseudo gap opens, the Hall coefficient was found to become strongly temperature-dependent. A suppression in resistivity was found in underdoped $`\text{(La, Sr)}_2\text{RuO}_4`$ as the pseudo gap opened. A similar feature has not been observed in $`\text{Sr}_2\text{RuO}_4`$. As mentioned above, the pseudo gap in $`\text{Sr}_2\text{RuO}_4`$ is present only in a small portion of the Fermi surface. The magnitude of the gap is also small as compared with that of the high-$`T_c`$ cuprates, resulting in unobservable effects in $`\rho _{ab}`$ and $`\rho _c`$. Finally we briefly mention our $`\mathrm{\Delta }\rho _c/\rho _c`$ results which will be discussed in detail in a future publication. As mentioned above, previously published work on $`\mathrm{\Delta }\rho _c/\rho _c`$ of $`\text{Sr}_2\text{RuO}_4`$ was focused on relatively high-temperature ($`>`$22 K) behavior although data obtained at 3.6 K were also shown. The magnitude of our longitudinal $`c`$-axis MR, $`\mathrm{\Delta }\rho _c^{}/\rho _c`$, is comparable with that of the transverse MR, $`\mathrm{\Delta }\rho _c^{}/\rho _c`$ . The difference, $`\mathrm{\Delta }\rho _c^{}/\rho _c`$ \- $`\mathrm{\Delta }\rho _c^{}/\rho _c`$, the ”pure” orbital MR, has very good $`H^2`$-dependence. A sign reversal from positive at low $`T`$ to negative at high $`T`$ was found in both $`\mathrm{\Delta }\rho _c^{}/\rho _c`$ and $`\mathrm{\Delta }\rho _c^{}/\rho _c`$ around 75 K, consistent with previous observations. However, different from the previous study, we found that $`\mathrm{\Delta }\rho _c(H)/\rho _c=\alpha H^2\beta H^4`$ (where $`\alpha `$ and $`\beta `$ are constants), actually describes well our $`\mathrm{\Delta }\rho _c^{}/\rho _c`$ and $`\mathrm{\Delta }\rho _c^{}/\rho _c`$ data. In summary, we have studied the magnetic field dependence of in-plane MR of $`\text{Sr}_2\text{RuO}_4`$ up to 7.3T. The linear positive transverse and longitudinal in-plane MR, not expected from band theory, has been found above a threshold field at low temperatures. We argue that such behavior is a manifestation of a coherent pseudo gap state in $`\text{Sr}_2\text{RuO}_4`$ formed at low temperatures. Theoretical input is needed to fully understand the experimental results. We would like to acknowledge helpful discussions with P. Coleman, N. Hussey, T. Imai, J. Jain, H. Kawano, Y. B. Kim, and G. Lonzarich, technical assistance from Yu. Zadorozhny, and finally many important contributions of D. Schlom to this project. This work is supported in part in the US by NSF through grants DMR-9702661 and ECS-9705839 and by the BMBF (Project number 13N6918/1) in Germany.
no-problem/9908/cond-mat9908304.html
ar5iv
text
# Quasiparticle Interactions in Fractional Quantum Hall Systems: Justification of Different Hierarchy Schemes ## I Introduction The fractional quantum Hall (FQH) effect, i.e. quantization of the Hall conductance of a two dimensional electron gas (2DEG) at certain densities in high magnetic fields, is a consequence of a finite excitation gaps that open above (incompressible) non-degenerate ground states (GS’s) at certain fractional fillings of the lowest Landau level (LL), $`\nu =1/3`$, 2/5, etc. Because of the LL degeneracy, it is clear that the those incompressible GS’s must originate from electron-electron interactions. This makes physics of the FQH effect very different from that underlying the integer quantum Hall (IQH) effect, despite similar manifestation of the two in transport experiments. While the origin of most prominent FQH states at $`\nu =1/m`$ ($`m`$ is an odd integer) has been explained by Laughlin, our understanding of why such states are also formed at some other filling factors (independent of the sample, density, etc.) or what conditions must in general be satisfied for an interacting system to exhibit the FQH effect at different fractions, is not yet completely satisfactory. As an extension of Laughlin’s idea, Haldane, and others proposed different, although equivalent, hierarchy schemes, in which the elementary quasiparticle-like excitations of the Laughlin fluid form Laughlin like states of their own. However, the original hierarchy approaches all share a major problem: they predict too many fractions and give no apparent connection between the stability of a given state and the hierarchy level at which it occurs. It is known that the FQH effect does not occur at some simple fractions predicted at low levels of hierarchy, while explanation of some other, experimentally observed ones requires introducing many generations of excitations. A different approach, introduced by Jain, and developed by Lopez and Fradkin, and by Halperin, Lee, and Read involves the concept of composite Fermions (CF’s). As formally described by the mean field Chern-Simons (CS) transformation, the CF’s are constructed by binding part of the external magnetic field $`B`$ to electrons in form of infinitely thin solenoids each carrying an even number $`2p`$ of flux quanta. The resulting CF’s move in an effective, reduced field $`B^{}`$. Smaller LL degeneracy of CF’s leads to a lower effective filling factor $`\nu ^{}`$ given by $$(\nu ^{})^1=\nu ^12p.$$ (1) It was found that the sequence of fractional electron filling factors $`\nu `$ corresponding to integer CF fillings $`\nu ^{}`$ agrees with the values at which FQH states occur. The qualitative success of the CF picture led to the conjecture that the CS charge-flux and Coulomb charge-charge interactions beyond the mean field largely cancel one another, and the system of strongly interacting electrons in a high magnetic field is converted to one of weakly interacting CF’s in a lower field. However, this cannot possibly be correct, since the CS energy ($`B`$) and Coulomb energy ($`\sqrt{B}`$) scale differently with the magnetic field $`B`$. Also, higher LL’s used in the procedure to obtain the wavefunctions are later eliminated by projection, which puts in doubt their role in the final result and most likely does not allow for interpretation in terms of an additional, effective magnetic field. In this paper we attempt to justify Haldane’s hierarchy picture in terms of the behavior of the interaction between quasiparticles (QP’s) which he considered of importance but totally unknown. We derive simple rules for identifying fractions that do or do not correspond to incompressible FQH states by analyzing pseudopotentials of the QP interactions on successive levels of hierarchy. Based on the connection between the form of pseudopotential (pair energy vs. pair angular momentum) and the nature of low lying states, we explain why QP’s do not form Laughlin states at all $`1/m`$ fillings. Our results validate understanding of experimentally observed FQH states in intuitive terms of the hierarchy of Laughlin excitations, established in Refs. . We also show that (in large systems) valid incompressible FQH states obtained in this hierarchy picture are equivalent to Jain states, despite different underlying physics used to justify the two approaches. This explains the success of the CF picture when applied to FQH systems better than assumed cancellation between CS and Coulomb interactions. It also defines limitations of the CF picture when applied to systems with general interactions (e.g., recently studied FQH systems of charged excitonic complexes formed in an electron-hole gas in a high magnetic field). The discussion is illustrated with results of exact numerical diagonalization calculations on a Haldane sphere of up to twelve electrons at $`\nu 1/3`$, (hamiltonians with dimensions of up to $`2\times 10^6`$) carried up using a modified Lanczos algorithm. ## II Haldane sphere Because in the absence of electron-electron interactions all many body states in the lowest LL ($`\nu 1`$) are degenerate, those interactions cannot be treated perturbatively. Therefore, numerical diagonalization techniques are commonly used to study FQH systems, which, however, limit their size to a finite (small) number of electrons. In order to model a finite density 2DEG, motion of $`N`$ electrons must be restricted to a finite area. This can be accomplished by imposing lateral confinement, periodic boundary conditions (PBC), or confining electrons on a closed surface (Haldane sphere). The last approach is particularly useful as it naturally avoids edge effects and preserves full 2D translational symmetry of a planar 2DEG in form of the rotational symmetry of a sphere. A pair of good quantum numbers on a plane, the center of mass (CM) and relative momenta, correspond to the total angular momentum $`L`$ and its projection $`L_z`$ on a sphere. The degeneracies associated with CM excitations on a plane correspond to those of $`L`$ multiplets on a sphere, and the non-degenerate GS’s of a planar 2DEG have $`L=0`$ in their spherical models. The magnetic field $`B`$ perpendicular to the surface of the Haldane sphere of radius $`R`$ is an isotropic radial field produced by a magnetic monopole placed at the origin. The monopole strength $`2S`$, defined as the number of flux quanta piercing the sphere, is an integer, as required by Dirac’s condition. In extrapolation of finite size results to the thermodynamic limit, the magnetic length $`\lambda =R/\sqrt{S}`$ is used as the length scale. The single particle states on the Haldane sphere are labeled by angular momentum $`lS`$ and its projection $`l_z`$, and are called monopole harmonics. The energies form $`(2l+1)`$-fold degenerate angular momentum shells, or LL’s, labeled by $`n=lS`$ and separated by cyclotron gaps. For the FQH states at $`\nu <1`$, only the lowest ($`n=0`$) spin polarized angular momentum shell of $`l=S`$ need be considered. The $`N`$ electron Laughlin $`\nu =1/m`$ states in a $`2l+1`$ degenerate shell occur at $`2l=m(N1)`$. ## III Hierarchy of Laughlin excitations The Haldane’s hierarchy of FQH states is constructed in the following way. At certain filling factors, $`\nu =1/m`$ ($`m`$ is an odd integer), electrons form Laughlin incompressible states. At $`\nu `$ slightly different from a Laughlin $`1/m`$ filling, the low lying states must contain a number of quasiparticles (QP’s) — quasielectrons (QE’s) at $`\nu >1/m`$ or quasiholes (QH’s) at $`\nu <1/m`$ — in the Laughlin $`\nu =1/m`$ state. States involving more than the minimum number of QP’s required by the difference between $`\nu `$ and $`1/m`$ contain additional QE-QH pairs and are separated from the lowest band by a gap $`\mathrm{\Delta }`$. The QP’s in the underlying (parent) Laughlin state have certain single particle energy $`\epsilon _{\mathrm{QP}}`$, statistics, number of available single particle states $`g_{\mathrm{QP}}`$ (in analogy to LL degeneracy for electrons), and interact with one another. Provided their interaction is small compared to $`\mathrm{\Delta }`$ and $`\epsilon _{\mathrm{QP}}`$, QP’s can, in principle, form Laughlin incompressible (daughter) states of their own. If $`\nu `$ is not exactly equal to the value at which QP’s would form a daughter Laughlin state, low lying QP states will contain their own QP-like excitations in the nearest daughter Laughlin state, which can in turn form an incompressible granddaughter Laughlin state, etc. Since the QE or QH statistics enters the hierarchy of fractions only through the counting of many body states, different statistics give equivalent results as long as the single particle degeneracies $`g_{\mathrm{QE}}`$ and $`g_{\mathrm{QH}}`$ are chosen correctly. The mean field CS transformation can be used to formally convert Bosons into Fermions by attaching one flux quantum to each Boson. For example, in the spherical geometry, where $`g_{\mathrm{QP}}=2l_{\mathrm{QP}}+1`$ is related to QP angular momentum $`l_{\mathrm{QP}}`$, systems of $`N`$ Bosons each with angular momentum $`k`$ and $`N`$ Fermions each with angular momentum $`l`$ are equivalent and contain the same $`L`$ multiplets if $$l=k+(N1)/2.$$ (2) As shown by Haldane (and can be understood from a simple picture of excitations created between electrons on a line), QE’s and QH’s in an $`N`$ electron Laughlin state can be viewed as Bosons with $`k_{\mathrm{QE}}=k_{\mathrm{QH}}=N/2`$. However, this combination of statistics and degeneracy gives proper counting of many body states only if an additional hard core is introduced that forbids two QE’s to be in a pair state with $`L=N`$. Such hard core can be accounted for by a mean field Fermion-to-Fermion CS transformation that replaces $`k_{\mathrm{QE}}`$ by $`k_{\mathrm{QE}}^{}=k_{\mathrm{QE}}(N_{\mathrm{QE}}1)`$, where $`N_{\mathrm{QP}}`$ is the QP number. In order to stress the connection with Jain’s CF picture and the recently proposed hierarchy of CF excitations, we use here a fermionic description of QP’s. The appropriate QP angular momenta obtained from Eq. (2) are $$l_{\mathrm{QP}}=\frac{N\pm (n_{\mathrm{QP}}1)}{2},$$ (3) with $`()`$ for $`l_{\mathrm{QE}}`$ and $`(+)`$ for $`l_{\mathrm{QH}}`$. Note that $`l_{\mathrm{QE}}`$ and $`l_{\mathrm{QH}}`$ given by Eq. (3) are equal to angular momenta of holes in highest filled and particles in lowest empty CF LL’s used in the CF picture. The expression for the filling factor $`\nu `$ of a daughter state is very similar to Eq. (1) and reads $$\nu ^1=2p+(1\pm \nu _{\mathrm{QP}})^1,$$ (4) where $`\nu _{\mathrm{QP}}`$ is the filling factor of QE’s $`(+)`$ or QH’s $`()`$ in Laughlin $`(2p+1)^1`$ parent state. Iteration of Eq. (4) with $`\nu _{\mathrm{QP}}`$ substituted by $`\nu `$ gives a continuous fraction, terminated when an incompressible state (without QP’s, $`\nu _{\mathrm{QP}}=0`$) is reached. E.g., for the state containing $`\nu =1/5`$ filling of QH’s in the $`\nu =1`$ filling of QE’s in the Laughlin $`\nu =1/3`$ state of electrons, this procedure gives: $$\nu =\frac{1}{21+{\displaystyle \frac{1}{1+{\displaystyle \frac{1}{20+{\displaystyle \frac{1}{1{\displaystyle \frac{1}{22+{\displaystyle \frac{1}{1+0}}}}}}}}}}}=\frac{9}{23}.$$ (5) In Fig. 1 we display more filling factors obtained in the hierarchy scheme. The lines connect parent states with their daughter states obtained for QP fillings $`\nu _{\mathrm{QP}}`$ shown in boxes. Note that holes created in a parent state of holes are particles in the grandparent state (in a sense that their presence increases overall $`\nu `$, i.e. their number increases when the magnetic field is decreased). E.g., holes created in $`\nu =1/3`$ filling of holes in $`\nu =1/3`$ Laughlin state of electrons will be referred to as QE’s in the $`\nu =2/7`$ state. As first stated by Haldane, whether a given fraction obtained in the hierarchy scheme corresponds to a stable incompressible GS depends on the stability of the parent state and on the interaction between QP’s in the daughter state. As will be shown in the following sections, these criteria eliminate most of all odd denominator fractions that can be constructed by an iteration of Eq. (4). The relatively small number of possible candidates left include the Jain sequence obtained from Eq. (1) for integer $`\nu ^{}`$. These states (and their electron-hole conjugates) have been marked with asterisks in Fig. 1. As will be shown in the following sections, these are all the incompressible states predicted by the “correct” hierarchy picture in the thermodynamic limit, and all of them have been confirmed experimentally. The fractions with a question mark in Fig. 1 are most likely compressible in the thermodynamic limit, but valid non-degenerate ($`L=0`$) hierarchy GS’s can occur at these fillings in finite systems. Those finite size valid hierarchy GS’s (Laughlin states of Laughlin QP’s) should be distinguished from other $`L=0`$ GS’s that can occur at different combinations of $`N`$ and $`2S`$ (e.g., for $`N=12`$ and $`2S=29`$ in Fig. 3c of Ref. ), but do not have Laughlin like QP correlations and thus cannot be associated with a filling factor. At the remaining filling factors, unmarked in Fig. 1, the hierarchy picture fails as expected and the system is compressible. ## IV Pseudopotentials The two body interaction of identical particles in an angular momentum shell of degeneracy $`2l+1`$ can be written in terms of the pseudopotential $`V()`$, i.e. the pair interaction energy $`V`$ as a function of relative pair angular momentum $$=2lL,$$ (6) where $`L`$ is the total angular momentum of the pair. $``$ is an odd integer and increases with increasing average separation. Plotting $`V`$ as a function of $``$ rather than of $`L`$ allows for meaningful comparison of pseudopotentials in shells of different degeneracy; for $`l\mathrm{}`$, the pseudopotentials calculated on the sphere converge to the pseudopotential on a plane (on a plane, $``$ is defined as the usual relative angular momentum). Whether a system of interacting Fermions will form a Laughlin state at the $`1/m`$ filling of their angular momentum shell depends on the short range of repulsive interaction. Precisely, the Laughlin $`\nu =1/m`$ incompressible state (in which $`m`$ for all pairs) is formed if, in the vicinity of $`=m`$, the interaction pseudopotential increases more quickly than linearly as a function of $`L(L+1)`$, i.e. more quickly than that of harmonic repulsion, $`V_\mathrm{H}(L)=\alpha +\beta L(L+1)`$, where $`\alpha `$ and $`\beta `$ are constants. More generally, if this condition is satisfied in the vicinity of $`=2p+1`$ for certain $`p`$, the total many body Hilbert space $``$ contains an (approximate) eigensubspace $`_p`$ holding states with $`2p+1`$ for all pairs (i.e. avoiding $`p`$ pair states of largest repulsion). A corresponding low energy band occurs in the spectrum, separated from higher states by a gap associated with $`V()`$. At Laughlin fillings of $`\nu =(2p+1)^1`$, the subspace $`_p`$ contains a single non-degenerate ($`L=0`$) multiplet with $`2p+1`$, and the lowest band consists of the Laughlin GS. The mathematical formalism derived to quantitatively treat the ability of electrons to avoid certain pair states involves the concept of fractional (grand)parentage, well known in atomic and nuclear physics and used recently to describe FQH systems. It is worth noting that avoiding highest energy states of three or more particles was recently proposed to explain incompressible GS’s at other fillings than Laughlin’s $`\nu =1/m`$. The electron (Coulomb) pseudopotential in the lowest LL $`V_\mathrm{e}()`$ satisfies the “short range” criterion (i.e. increases more quickly than $`V_\mathrm{H}`$ as a function of $`L`$) in entire range of $``$, which is the reason for incompressibility of principal Laughlin $`\nu =1/m`$ states. However, this does not generally hold for the QP pseudopotentials on higher levels of hierarchy. We have obtained some of those pseudopotentials for different values of $`l_{\mathrm{QP}}`$ by numerical diagonalization of appropriate many electron hamiltonians on the Haldane sphere and identification of the lowest bands in obtained energy spectra. The total many electron energies within those bands contain the energy $`E_0`$ of the parent state, single particle energies $`2\epsilon _{\mathrm{QP}}`$ of the pair of appropriate QP’s, and the QP-QP interaction energy $`V_{\mathrm{QP}}(L)`$. In Fig. 2 we show the results for QE’s and QH’s in Laughlin $`\nu =1/3`$ (data for $`N8`$ was published before in Ref. ) and $`\nu =1/5`$ states. The plotted energy $`E()=2\epsilon _{\mathrm{QP}}+V_{\mathrm{QP}}(L)`$ is given in units of $`e^2/\lambda `$ where $`\lambda `$ is the magnetic length in the parent state. Different symbols mark pseudopotentials obtained in diagonalization of $`N`$ electron systems with different $`N`$ and thus with different $`l_{\mathrm{QP}}`$, see Eq. (3). Clearly, the QE and QH pseudopotentials are quite different and neither one decreases monotonically with increasing $``$. On the other hand, the corresponding pseudopotentials in $`\nu =1/3`$ and 1/5 states look similar, only the energy scale is different. The convergence of energies at small $``$ obtained for larger $`N`$ suggests that the maxima at $`=3`$ for QE’s and at $`=1`$ and 5 for QH’s, as well as the minima at $`=1`$ and 5 for QE’s and at $`=3`$ and 7 for QH’s, persist in the limit of large $`N`$ (i.e. for an infinite system on a plane). Consequently, the only incompressible daughter states of Laughlin $`\nu =1/3`$ and 1/5 states are those with $`\nu _{\mathrm{QE}}=1`$ or $`\nu _{\mathrm{QH}}=1/3`$ (asterisks in Fig. 1) and (maybe) $`\nu _{\mathrm{QE}}=1/5`$ and $`\nu _{\mathrm{QH}}=1/7`$ (question marks in Fig. 1). It is also clear that no incompressible daughter states will form at e.g. $`\nu =4/11`$ or 4/13. Let us note that the incompressibility of daughter states with completely filled QE shell (e.g., at $`\nu =2/5`$ or 2/9) does not require any special form of the QE-QE interaction, except that it must be weaker than the single particle energies $`\epsilon _{\mathrm{QE}}`$ and $`\epsilon _{\mathrm{QH}}`$ responsible for the gap. In this sense, the FQH effect at Jain filling $`\nu =2/5`$ can be viewed as an IQH effect of QE’s in the Laughlin $`\nu =1/3`$ state, except that the degenerate single particle shell available to QE’s is due to special form of elementary excitations of the parent Laughlin state rather than due to an effective magnetic field. Similarly, the excitation gap at the $`\nu =2/5`$ filling is not a cyclotron gap but the energy needed to create a QE-QH pair like excitation in the filled shell of Laughlin QE’s. On the other hand, the FQH effect at $`\nu =2/7`$ is a fractional effect also on the level of QH’s in the parent $`\nu =1/3`$ state, and its excitation gap is governed by $`V_{\mathrm{QH}}(1)`$, the largest pseudopotential parameter for QH-QH interaction. The electron pseudopotential $`V_\mathrm{e}()`$ is not strictly a short range one (for which $`V(1)V(3)\mathrm{}`$) and the associated hidden symmetry responsible for occurrence of eigensubspaces $`_p`$ and incompressible Laughlin states is only approximate. Actually, a fairly small reduction of $`V_\mathrm{e}(1)`$ compared to $`V_\mathrm{e}(3)`$ that can be achieved in a wide quantum well leads to a break down of Laughlin $`\nu =1/3`$ state. While the hidden symmetry makes the low lying states near Laughlin fillings virtually insensitive to the details of $`V_\mathrm{e}()`$ as long as it has short range (i.e. increases more quickly than $`V_\mathrm{H}`$ as a function of $`L`$), it is interesting to ask to what extent the form of $`V_\mathrm{e}()`$ affects the QP pseudopotentials, and thus the incompressibility of related daughter states. We have compared the pseudopotentials in Fig. 2 with the ones obtained for “exponential” interaction, $`V_{\mathrm{exp}}(+2)=V_{\mathrm{exp}}()/10`$, and for “selective” interaction, $`V_{\mathrm{sel}}(<m)=\mathrm{}`$, $`V_{\mathrm{sel}}(m)=1`$, and $`V_{\mathrm{sel}}(>m)=0`$, and found that all the features in Fig. 2 remain unchanged. This means that the short range character of interaction between particles in the parent state does not imply the same for interaction between Laughlin QP’s in the daughter state. This observation, essential for understanding why incompressible states do not occur at all odd denominator fractions, might appear somewhat surprising since the QP’s are (fractionally) charged objects and hence their interaction has similar nature to that of electrons. However, it must be kept in mind that it is the combination of interaction potential $`V(r)`$ and available single particle Hilbert space that gives pseudopotential $`V()`$ and, in turn, determines if Laughlin states occur (e.g., for the same Coulomb potential $`V(r)=e^2/r`$, Laughlin $`\nu =1/3`$ state occurs in the lowest LL but does not occur in the excited ones). Once the Laughlin like states on a given hierarchy level are found, the QP pseudopotentials must be calculated for these states to determine if they in turn can have any incompressible daughter states. As an example of this procedure, in Fig. 3 we present a few pseudopotentials calculated for the $`\nu =2/5`$, 2/7, and 2/9 parent states. As in Fig. 2, energy $`E()=2\epsilon _{\mathrm{QP}}+V_{\mathrm{QP}}(L)`$ is given in the units of $`e^2/\lambda `$, with $`\lambda `$ appropriate for the parent state, and $`N`$ is the number of electrons in the system that was diagonalized to obtain a particular pseudopotential. The pseudopotentials plotted in Fig. 3 show two types of behavior at small $``$. The ones in Fig. 3bcf have a maximum at $`=3`$, similarly as those in Fig. 2ac, while the ones in Fig. 3ade increase when $``$ increases between 1 and 5. Similar behavior of different pseudopotentials is a consequence of the particle-hole symmetry between QE’s of a parent state and QH’s of its daughter state with filled QE shell ($`\nu _{\mathrm{QE}}=1`$, see Fig. 1). E.g., vacancies in an almost completely filled shell of QE’s in the $`\nu =1/3`$ state are QH’s of the $`\nu =2/5`$ state; vacancies in an almost completely filled shell of QE’s in the $`\nu =2/5`$ state are QH’s of the $`\nu =3/7`$ state, etc. The relation between QH’s in the parent state and QE’s in the daughter state with $`\nu _{\mathrm{QH}}=1`$ is equivalent, as the latter is simply the grandparent state. The particle-hole symmetry discussed above is only approximate because the single particle gaps are not infinitely large compared to single particle gaps $`\mathrm{\Delta }`$ and $`\epsilon `$, and the QP number is not strictly conserved. However, the appropriate pseudopotentials are to a good approximation equal, and e.g. by comparing our data for QE’s in $`\nu =1/3`$ and QH’s in $`\nu =2/5`$ we were able to extract energies $`\epsilon _{\mathrm{QH}}`$ of a single QH in the $`\nu =2/5`$ state as a function of the system size $`N`$. The linear extrapolation to $`1/N0`$ gives $`\epsilon _{\mathrm{QH}}=0.0098`$ $`e^2/\lambda `$, and the limiting value of the “proper” QH energy (including additional fractional charge $`e/5`$ in the background) is $`\stackrel{~}{\epsilon }_{\mathrm{QH}}=0.0123`$ $`e^2/\lambda `$. It is apparent from Fig. 3 that the incompressible daughter states derived from $`\nu =2/5`$, 2/7, or 2/9 must either have a completely filled QE shell (Jain fractions $`\nu =3/7`$, 3/11, and 3/13, respectively), or (possibly) the 1/5 filling of QH’s ($`\nu =9/23`$, 9/31, and 9/41, respectively). Incompressible daughter states at any other fractions, including the 1/3 filling of QE’s or QH’s (giving such fractions as $`\nu =5/13`$ or 7/17, see Fig. 1) do not occur. ## V Numerical tests for finite systems In order to test the predictions of low lying states in terms of Laughlin QP’s interacting through appropriate pseudopotentials, we have calculated numerically exact energy spectra of up to twelve electrons on the Haldane sphere at different values of the monopole strength $`2S`$, i.e., different filling factors. As demonstrated on the examples presented in Fig. 4, the results (in all cases we looked at) can be very well understood in terms of QP-QP interaction. Fig. 4a shows the spectrum of eight electrons at $`2S=18`$. The low lying band contains states of three QE’s in Laughlin $`\nu =1/3`$ state, each with $`l_{\mathrm{QE}}=3`$ ($`L=0`$, 2, 3, 4, and 6). As marked with a dashed line, there is a gap separating the low energy band from higher states. Due to the QE-QE interaction, the lowest band is not degenerate and has certain width. Because this width is small compared to the gap to higher states (i.e. to energy $`\mathrm{\Delta }`$ to create additional QE-QH pairs), the three QE’s interact with one another in the presence of a rigid background (Laughlin fluid at $`\nu =1/3`$) and the low lying states are determined by the pseudopotential $`V_{\mathrm{QE}}()`$ obtained for the same $`l_{\mathrm{QE}}=3`$ ($`N=7`$ in Fig. 2a). If $`V_{\mathrm{QE}}()`$ had short range, the multiplet at $`L=0`$ would be an incompressible GS corresponding to the Laughlin $`\nu _{\mathrm{QE}}=1/3`$ state and total electron filling factor of $`\nu =4/11`$. However, as discussed in the previous section, $`V_{\mathrm{QE}}()`$ has a minimum at $`=1`$ and the system is compressible (in this small system, the Laughlin $`\nu _{\mathrm{QE}}=1/3`$ state is an eigenstate as it is the only state of three QE’s in the $`L=0`$ subspace; in larger systems, it will mix with other $`L=0`$ states and fall into the continuum). A similar spectrum is displayed in Fig. 4b. Here, the low lying states contain three QH’s in the Laughlin $`\nu =1/3`$ state, each with $`l_{\mathrm{QH}}=5`$. As expected from the discussion of $`V_{\mathrm{QH}}`$, the states with $`3`$ ($`L=0`$, 2, 3, 4, and 6) have lowest energy within this band, but the Laughlin $`\nu _{\mathrm{QH}}=1/5`$ state with $`5`$ in not the GS and the system at $`\nu =4/13`$ is compressible. Since the GS’s at $`\nu =4/11`$ or 4/13 are not valid parent states (QP Laughlin states), the analysis of states at $`\nu `$ near 4/11 or 4/13 in terms of their daughter QP’s is not possible. This is demonstrated in Fig. 4cd, where the states that would contain two QE’s in the $`\nu =4/11`$ each with $`l_{\mathrm{QE}}=3/2`$ ($`L=0`$ and 2) and two QE’s in the $`\nu =4/13`$ each with $`l_{\mathrm{QE}}=2`$ ($`L=1`$ and 3), do not have lowest energy. The low lying states in each of the spectra in Fig. 4ef contain three QP’s each with $`l_{\mathrm{QP}}=3`$ in a valid incompressible higher level hierarchy state ($`L=0`$, 2, 3, 4, and 6). The interaction of QE’s in the $`\nu =2/5`$ state and of QH’s in the $`\nu =3/7`$ state is similar as those particles are connected through the particle-hole symmetry. Also, since the appropriate pseudopotential in Fig. 3a (for $`N=10`$) increases with increasing $``$, we expect the standard atomic Hund rule to hold, i.e. the lowest energy state should have maximum allowed $`L`$ within the lowest energy band of three QP’s. Indeed, the $`L=6`$ state is the GS of both systems, while the states at $`L=0`$ predicted by the hierarchy picture ($`\nu =7/17`$ and 8/19) have higher energy. Numerical tests of other fractions become difficult due to increasing size of the system. However, one can diagonalize the hamiltonian of interacting QP’s in a given parent states using their known single particle energies $`\epsilon _{\mathrm{QP}}`$ (the values for QP’s in Laughlin $`\nu =1/3`$ and 1/5 states can be found in Refs. ) and pseudopotentials $`V_{\mathrm{QP}}()`$, and obtain approximate lowest energy levels of an underlying (larger) electron system (with respect to the energy $`E_0`$ of the parent incompressible state). The error made in such approximate calculation is due to neglected scattering processes involving other objects than the specified QP’s (e.g. polarization of the parent state through creation of additional QE-QH pairs, etc.). At least for states with largest $`\epsilon _{\mathrm{QP}}`$ (compared to the strength of QP interactions), this error is expected to be small, which validates the tests of incompressibility of their daughter states. The results of such tests for a few different systems are shown in Fig. 5. In all frames, the energy is measured from the energy of the parent ($`\nu =1/3`$) state and $`\lambda `$ is the magnetic length in the parent state. In Fig. 5ab the approximate spectra (full dots) are overlaid with exact energies (pluses) obtained by diagonalization of the full electron hamiltonian. Clearly, both for interacting QE’s and QH’s, the approximate calculation gives the lowest lying states with negligible error (most of which is due to different magnetic lengths in the daughter and parent states and can be corrected; error due to neglected scattering processes is hardly visible). This agreement proves that the low lying states in FQH systems indeed contain QP’s characterized by certain level degeneracy and interaction (weakly dependent on the QP number), and validates use of the approximate calculation for larger systems. The data in Fig. 5a is by itself another example showing that QE’s in Laughlin $`\nu =1/3`$ state do not form a separate band of states with $`3`$ (these would be states with $`L=3/2`$, 5/2, and 9/2). Similarly, Fig. 5b shows that QH’s form a band with $`3`$ ($`L=3/2`$, 5/2, 7/2, 9/2, 11/2, and 15/2), but do not form a band with $`5`$ (here, one multiplet with $`L=3/2`$ that would correspond to a QE in the $`\nu =4/13`$ state). Fig. 5cdef shows approximate spectra of larger systems. Fig. 5c shows no band corresponding to two QE’s in the $`\nu =4/11`$ state of fourteen electrons (these would be states at $`L=1`$ and 3), Fig. 5d shows a band corresponding to two QE’s in the $`\nu =2/7`$ state of ten electrons ($`L=1`$, 3, and 5; this is the QE pseudopotential like those for $`N=6`$ and 8 in Fig. 3c), Fig. 5e shows no band of (Laughlin $`\nu =1/3`$) QE states with $`3`$ (these would be states at $`L=0`$, 2, 3, 4, and 6), and an incompressible $`\nu =6/17`$ state of twelve electrons corresponding to $`\nu _{\mathrm{QE}}=1/5`$, and Fig. 5f shows a band of (Laughlin $`\nu =1/3`$) QH states with $`3`$, no band with $`5`$ (these would be states at $`L=0`$, 2, 3, 4, and 6), and an incompressible $`\nu =6/19`$ state of twelve electrons corresponding to $`\nu _{\mathrm{QE}}=1/7`$. The $`\nu =6/17`$ and 6/19 states of twelve electrons in Fig. 5ef are the only non-Jain hierarchy states we have tested numerically which are predicted to be (incompressible) $`L=0`$ GS’s based on pseudopotentials in Figs. 2 and 3 and which contain at least three QP’s. However, because of the particle-hole symmetry and the similarity of different QP pseudopotentials (see Figs. 2 and 3), the numerical evidence for valid $`L=0`$ hierarchy ground states at $`\nu =6/17`$ and 6/19 (in a twelve electron system) suggests stability of some other hierarchy states, e.g. at $`\nu =6/29`$, 6/31, 9/23, an 11/39 (all states with question marks in Fig. 1), at least in finite systems with appropriate electron number (see next section). ## VI Predictions for infinite systems The extrapolation of our numerical results in order to predict stability of different hierarchical states in infinite systems must be done very carefully. The calculations show that the interaction of QP’s in a Laughlin parent state is not generally repulsive. The pseudopotential $`V_{\mathrm{QP}}()`$ is obtained by subtraction of two appropriate QP energies $`2\epsilon _{\mathrm{QP}}`$ from energies $`E()`$ in Figs. 2 and 3. E.g., for the $`\nu =1/3`$ parent state, the only positive pseudopotential parameter is $`V_{\mathrm{QH}}(1)`$, and all others ($`V_{\mathrm{QH}}()`$ for $`3`$ and $`V_{\mathrm{QE}}()`$ for all $``$) are negative. Since $`V_{\mathrm{QP}}()`$ at large $``$ (large distance) is expected to vanish, it must also increase above all the values at small $``$ except for $`V_{\mathrm{QH}}(1)`$. This brings out the question if our prediction of incompressible states at e.g. $`\nu =6/17`$ ($`\nu _{\mathrm{QE}}=1/5`$) or $`\nu =6/19`$ ($`\nu _{\mathrm{QH}}=1/7`$), verified numerically for twelve electrons (Fig. 5ef) remains valid for an infinite system. We have calculated energy spectra of six particles at filling $`\nu =1/5`$ ($`2l=25`$), interacting through pseudopotentials which for $`9`$ were equal to $`V_{\mathrm{QE}}()`$ for $`N=11`$ in Fig. 2a, vanished at $`=25`$, and behaved in a number of different ways for $`9<<25`$. As shown in Fig. 6, the $`L=0`$ hierarchy state is always the lowest energy state in the low $`L`$ part of the spectrum, but whether it is an absolute GS depends on how quickly $`V_{\mathrm{QP}}()`$ goes to zero with increasing $``$. This is a consequence of the following operator identity $$\underset{i<j}{}\widehat{L}_{ij}^2=\widehat{L}^2+N(N2)\widehat{l}^2,$$ (7) which relates total ($`L`$) and pair ($`\widehat{L}_{ij}`$) angular momenta of a system of $`N`$ particles in a shell of angular momentum $`l`$. The states with larger $`L`$ have (on the average) larger values of $`L_{ij}`$ and thus, if $`V()`$ increases with increasing $``$, lower energy. The exact numerical calculation of $`V_{\mathrm{QP}}()`$ for $`9`$, i.e. for $`l_{\mathrm{QP}}5`$, seems impossible (calculation for $`l_{\mathrm{QE}}=5`$, i.e. $`N=11`$, already required diagonalization of a matrix with dimension nearly $`10^6`$), and thus the only valid test of stability of states like $`\nu =6/17`$ or 6/19 might be experiment. However, since $`V_{\mathrm{QP}}()`$ for large (infinite) system should be virtually zero for all $``$’s above certain critical value, it is unlikely that these GS will be incompressible in the thermodynamic limit. This would explain why (to our knowledge), no such states have been observed. The above arguments most likely eliminate all daughter states derived from Laughlin $`\nu =1/m`$ parents with $`\nu _{\mathrm{QE}}=1/5`$ and $`\nu _{\mathrm{QH}}=1/7`$ as possible incompressible GS’s, leaving only those with $`\nu _{\mathrm{QE}}=1`$ and $`\nu _{\mathrm{QH}}=1/3`$. The latter ones are incompressible, because they either correspond to a filled shell ($`\nu _{\mathrm{QE}}=1`$) or avoid the (only) repulsive pseudopotential parameter $`V_{\mathrm{QH}}(1)>0`$ ($`\nu _{\mathrm{QH}}=1/3`$). A quick look at the pseudopotentials in Fig. 3 is enough to find that the incompressible states on higher levels of hierarchy can only have $`\nu _{\mathrm{QE}}=1`$. It is easy to check (see also Fig. 1) that fractions generated in this way belong to the Jain sequence obtained from Eq. (1) for integer $`\nu ^{}`$. This explains why the hierarchy fractions from outside this sequence have not been observed experimentally, even though numerical results for finite systems presented here might suggest that states like $`\nu =6/17`$ or 6/19 are incompressible. Let us stress that the fact that valid hierarchy states (i.e. states with Laughlin like correlations of appropriate Laughlin QP’s) occur only at the Jain sequence of filling factors obviously does not contradict occurrence of FQH effect at other fractions. However, the correlations in other possible FQH states, the origin of their incompressibility, and their elementary excitations must be different. Examples of observed non-hierarchy FQH states include ones at $`\nu =7/3`$ (calculations show that it is not a Laughlin like $`\nu =1/3`$ state in the first excited LL) and $`\nu =5/2`$. ## VII Comparison with CF picture The mean field CF picture correctly predicts not only the incompressible states at the Jain sequence of filling factors, but also the low lying bands of states at any value of $`N`$ and $`2S`$. However, neither its original justification based on cancellation between Coulomb and CS interactions beyond the mean field nor the use of higher LL’s in construction of CF wavefunctions can be accepted as complete understanding of this success. If the effective CF magnetic field is non-negative ($`2S^{}0`$) and the effective CF filling factor is less or equal than one ($`\nu ^{}1`$) the CF picture selects out of the total Hilbert space $``$ the subspace $`_p`$, where $`2p`$ is the number of bound (attached) flux quanta. From Eq. (4) and Fig. 1 it is clear that entire sequence of Jain fractions corresponds to valid hierarchy states, obtained by for the QE filling of $`\nu _{\mathrm{QE}}=0`$ or 1 on any level of hierarchy and/or the QH filling of $`\nu _{\mathrm{QH}}=1/3`$ on the first level (QH’s in the principal Laughlin $`\nu =1/m`$ state). Note that CF states obtained with $`2p`$ bound flux quanta for which $`2S^{}`$ is positive are derived from the $`\nu _{\mathrm{QE}}=1`$ daughter of the Laughlin $`\nu =(2p+1)^1`$ state, while those for which $`2S^{}`$ is negative come from the $`\nu _{\mathrm{QH}}=1/3`$ daughter of the Laughlin $`\nu =(2p1)^1`$ state. The explicit hierarchy wavefunctions can also be constructed without introducing higher LL’s excitations. As demonstrated for few electron systems, the (valid) hierarchy and CF wavefunctions are nearly identical. Another qualitative success of the CF picture, the description of higher bands in the energy spectrum in terms of excitations between CF LL’s, in the hierarchy picture corresponds to creation of additional QE-QH pairs in the parent state. One of the main results used as a direct experimental evidence for the existence of CF’s – the observation of geometric resonances and divergence of the CF cyclotron radius at $`B^{}0`$ – does not contradict the hierarchy picture where the fractionally charged relevant QP’s move in a bare external field $`B`$ so that the cyclotron radius coincides with the one of CF’s moving in an effective field $`B^{}`$. Another result seemingly proving the formation of CF LL’s is the linear dependence of the excitation gap of Jain states on the effective magnetic field $`B^{}`$. In Fig. 7 we plot the gaps $`\mathrm{\Delta }_N`$ calculated numerically for a few most prominent hierarchy/Jain states as a function of the inverse electron number, $`1/N`$. The limiting values $`\mathrm{\Delta }_{\mathrm{}}`$ are plotted in the inset as a function of $`B^{}/B=12p\nu `$. The gaps of states obtained from Eq. (1) for $`p=1`$ ($`\nu =1/3`$, 2/5, 3/7, …) fall on a straight line vs. $`B^{}/B`$ (as first observed by Du et al.; note however that the linear extrapolation to $`B^{}=0`$ gives negative gaps, also in agreement with experiment of Du at al.). However, it is not so for Jain $`p=2`$ states at $`\nu =1/5`$, 2/7, and 2/9. In particular, the gap of the $`\nu =2/7`$ state seems to be larger than that of the $`\nu =1/5`$ state. While this result may be difficult to accept in the CF picture, it is by no means surprising in the hierarchy picture where the relevant QP’s in the two states are different and interact through different pseudopotentials (see Figs. 1, 2, and 3). The mean field CF picture and the present hierarchy picture are equivalent because they both use correct degeneracy of excitations, $`g_{\mathrm{QE}}`$ and $`g_{\mathrm{QH}}`$, for the chosen (fermionic) statistics. The CF picture makes no use of the form of single particle wavefunctions in excited LL’s; an effective magnetic field is just another way to obtain correct $`g_{\mathrm{QE}}`$ and $`g_{\mathrm{QH}}`$. However, the authors believe that the understanding of incompressible states at Jain filling factors in terms of hierarchy of Laughlin excitations and involved QP pseudopotentials has a number of advantages over the CF picture. It does not use such puzzling concepts as flux binding, depends explicitly and in a known way on the form of electron-electron interaction which enables predicting its applicability to systems with modified interactions (higher LL’s, finite well width, etc.), and predicts correct energy gaps in terms of interaction parameters rather than arbitrary effective cyclotron energy. The hierarchy picture is also more physically intuitive and makes the origin of incompressibility of Jain states more clear. Moreover, it gives better understanding of why no other fractions are experimentally observed, even though some (e.g. $`\nu =6/17`$ and 6/19) are found in finite size numerical calculations. ## VIII Conclusion We have calculated pseudopotentials of the interaction between quasiparticles, arising in the hierarchy picture of incompressible FQH states. Based on the analysis of these pseudopotentials, it is explained why no hierarchy states with filling factors $`\nu `$ from outside the Jain sequence occur in an infinite system, and thus why none have been observed experimentally. Compressibility of $`\nu =4/11`$ and 4/13 states is demonstrated. Laughlin like hierarchy states other than Jain states are found at $`\nu =6/17`$ and 6/19 in finite size numerical calculations. However, it is argued that the system at these fractions (and other non-Jain fractions with question marks in Fig. 1) will undergo a transition into a compressible phase when its size is increased. The descriptions of FQH states in terms of mean field CF’s and hierarchy of Laughlin excitations are compared. It is explained why, despite no rigorous justification of the CF assumption of flux (or vortex) binding, the CF predictions and the valid predictions of the hierarchy picture are (qualitatively) equivalent. In our analysis we used a fermionic statistics of quasiparticles. However, our results are independent of this statistics and remain valid for hierarchy pictures formulated in terms of bosons or anyons.
no-problem/9908/hep-ph9908360.html
ar5iv
text
# Mapping the QCD Phase Diagram ## 1 The Critical Point on the Map We begin our walk through the phase diagram at zero baryon number density, with a brief review of the phase changes which occur as a function of temperature. That is, we begin by restricting ourselves to the vertical axis in Figures 1 through 4. This slice of the phase diagram was explored by the early universe during the first tens of microseconds after the big bang and can be studied in lattice simulations. As heavy ion collisions are performed at higher and higher energies, they create plasmas with a lower and lower baryon number to entropy ratio and therefore explore regions of the phase diagram closer and closer to the vertical axis. In QCD with two massless quarks ($`m_{u,d}=0`$; $`m_s=\mathrm{}`$; Figure 1) the phase transition at which chiral symmetry is restored is likely second order and belongs to the universality class of $`O(4)`$ spin models in three dimensions . Below $`T_c`$, chiral symmetry is broken and there are three massless pions. At $`T=T_c`$, there are four massless degrees of freedom: the pions and the sigma. Above $`T=T_c`$, the pion and sigma correlation lengths are degenerate and finite. In nature, the light quarks are not massless. Because of this explicit chiral symmetry breaking, the second order phase transition is replaced by an analytical crossover: physics changes dramatically but smoothly in the crossover region, and no correlation length diverges. Thus, in Figure 2, there is no sharp boundary on the vertical axis separating the low temperature hadronic world from the high temperature quark-gluon plasma. This picture is consistent with present lattice simulations , which suggest $`T_c140170`$ MeV . Arguments based on a variety of models indicate that the transition as a function of $`T`$ is first order at large $`\mu `$. This suggests that the phase diagram features a critical point $`E`$ at which the line of first order phase transitions present for $`\mu >\mu _E`$ ends, as shown in Figure 2.<sup>1</sup><sup>1</sup>1If the up and down quarks were massless, $`E`$ would be a tricritical point , at which the first order transition becomes second order. See Figure 1. At $`\mu _E`$, the phase transition is second order and is in the Ising universality class . Although the pions remain massive, the correlation length in the $`\sigma `$ channel diverges due to universal long wavelength fluctuations of the order parameter. This results in characteristic signatures, analogues of critical opalescence in the sense that they are unique to collisions which freeze out near the critical point, which can be used to discover $`E`$ . Returning to the $`\mu =0`$ axis, universal arguments , again backed by lattice simulation , tell us that if the strange quark were as light as the up and down quarks, the transition would be first order, rather than a smooth crossover. This means that if one could dial the strange quark mass $`m_s`$, one would find a critical $`m_s^c`$ at which the transition as a function of temperature is second order . Figures 2, 3 and 4 are drawn for a sequence of decreasing strange quark masses. Somewhere between Figures 3 and 4, $`m_s`$ is decreased below $`m_s^c`$ and the transition on the vertical axis becomes first order. The value of $`m_s^c`$ is an open question, but lattice simulations suggest that it is about half the physical strange quark mass . These results are not yet conclusive but if they are correct then the phase diagram in nature is as shown in Figure 3, and the phase transition at low $`\mu `$ is a smooth crossover. These observations fit together in a simple and elegant fashion. If we could vary $`m_s`$, what we would find is that as $`m_s`$ is reduced from infinity to $`m_s^c`$, the critical point $`E`$ in the $`(T,\mu )`$ plane moves toward the $`\mu =0`$ axis . This is shown in Figures 2-4. In nature, $`E`$ is at some nonzero $`T_E`$ and $`\mu _E`$. When $`m_s`$ is reduced to $`m_s^c`$, between Figure 3 and Figure 4, $`\mu _E`$ reaches zero. Of course, experimentalists cannot vary $`m_s`$. They can, however, vary $`\mu `$. The AGS, with beam energy 11 AGeV corresponding to $`\sqrt{s}=5`$ GeV, creates fireballs which freeze out near $`\mu 500600`$ MeV . When the SPS runs with $`\sqrt{s}=17`$ GeV (beam energy 158 AGeV), it creates fireballs which freeze out near $`\mu 200`$ MeV . By dialing $`\sqrt{s}`$ and thus $`\mu `$, experimenters can find the critical point $`E`$. ## 2 Discovering the Critical Point The map of the QCD phase diagram which we have sketched so far is simple, coherent and consistent with all we know theoretically; the discovery of the critical point would provide an experimental foundation for the central qualitative feature of the landscape. This discovery would in addition confirm that in higher energy heavy ion collisions and in the big bang, the QCD phase transition is a smooth crossover. Furthermore, the discovery of collisions which create matter that freezes out near $`E`$ would imply that conditions above the transition existed prior to freezeout, and would thus make it much easier to interpret the results of other experiments which seek signatures which probe the partonic matter created early in the collision. We theorists must clearly do as much as we can to tell experimentalists where and how to find $`E`$. The “where” question, namely the question of predicting the value of $`\mu _E`$ and thus suggesting the $`\sqrt{s}`$ to use to find $`E`$, is much harder for us to answer. One of the things which is intrinsic to the picture we have described is that $`\mu _E`$ is sensitive to the mass of the strange quark, and therefore particularly hard to predict. Crude models suggest that $`\mu _E`$ could be $`600800`$ MeV in the absence of the strange quark ; this in turn suggests that in nature $`\mu _E`$ may have of order half this value, and may therefore be accessible at the SPS if the SPS runs with $`\sqrt{s}<17`$ GeV. However, at present theorists cannot predict the value of $`\mu _E`$ even to within a factor of two. The SPS can search a significant fraction of the parameter space; if it does not find $`E`$, it will then be up to the RHIC experiments to map the $`\mu _E<200`$ MeV region. The other half of the “where” question is the question of how close does one have to come to $`E`$ to detect its presence. That is, how big steps in $`\sqrt{s}`$ can one safely take and still be reasonably confident of discovering $`E`$ if it is there to be found? To answer this, we must estimate $`\mathrm{\Delta }\mu `$, the width in $`\mu `$ of the region centered at $`\mu _E`$ within which the correlation length is longer than $`2`$ fm, and thus detectable effects of $`E`$ arise . Here again, only crude models are available. Analysis within the toy model of Ref. suggests that in the absence of the strange quark, $`\mathrm{\Delta }\mu 120`$ MeV for $`\mu _E800`$ MeV. Similar results can be obtained within a random matrix model . It is likely over-optimistic to estimate $`\mathrm{\Delta }\mu 120`$ MeV when the effects of the strange quark are included and $`\mu _E`$ itself is significantly reduced. A conservative estimate would be to use the models to estimate that $`\mathrm{\Delta }\mu /\mu _E15\%`$ in an infinite system. Finite size effects must increase $`\mathrm{\Delta }\mu /\mu _E`$. As a reasonable estimate, the best available at present for use in planning experimental strategy, one may take $`\mathrm{\Delta }\mu /\mu _E2030\%`$. This suggests that if experiments could be done at about four energies between AGS energy ($`\sqrt{s}=5`$) and maximum SPS energy ($`\sqrt{s}=17`$ GeV), there would be a good chance of finding $`E`$ if it lies within the range $`200\mathrm{MeV}<\mu _E<600\mathrm{MeV}`$. An SPS run at $`\sqrt{s}=9`$ GeV (beam energy 40 AGeV) is already planned. This data, together with that already taken at $`\sqrt{s}=17`$ GeV, together with data from two (or perhaps only one) additional beam energies in between would allow the SPS to search for the critical point over a substantial range of parameter space. There is therefore a strong scientific argument for an 80 AGeV run. If runs at two energies between 40 AGeV and 158 AGeV were possible, that would be ideal. It should be clear by now that although we are trying to be helpful with the “where” question, we are not very good at answering it quantitatively. This question can only be answered convincingly by an experimental discovery. What we as theorists can do reasonably well is to answer the “how” question, thus enabling experimenters to answer “where”. This is the goal of a recent paper by Stephanov, myself and Shuryak . The signatures we have proposed are based on the fact that $`E`$ is a genuine thermodynamic singularity at which susceptibilities diverge and the order parameter fluctuates on long wavelengths. The resulting signatures are nonmonotonic as a function of $`\sqrt{s}`$: as this control parameter is varied, we should see the signatures strengthen and then weaken again as the critical point is approached and then passed. The simplest observables we analyze are the event-by-event fluctuations of the mean transverse momentum of the charged particles in an event, $`p_T`$, and of the total charged multiplicity in an event, $`N`$. We calculate the magnitude of the effects of critical fluctuations on these and other observables, making predictions which, we hope, will allow experiments to find $`E`$. As a necessary prelude, we analyze the contribution of noncritical thermodynamic fluctuations. We compare the noncritical fluctuations of an equilibriated resonance gas to the fluctuations measured by NA49 at $`\sqrt{s}=17`$ GeV . The observed fluctuations are as perfect Gaussians as the data statistics allow, as expected for freeze-out from a system in thermal equilibrium. The width of the event-by-event distribution of the mean $`p_T`$ is in good agreement with predictions based on noncritical thermodynamic fluctuations.<sup>2</sup><sup>2</sup>2This width can be measured even if one observes only two pions per event ; large acceptance data as from NA49 is required in order to learn that the distribution is Gaussian, that thermodynamic predictions may be valid, and that the width is therefore the only interesting quantity to measure. The data on multiplicity fluctuations show evidence for a nonthermodynamic contribution, which is to be expected since the extensive quantity $`N`$ is sensitive to the initial size of the system and thus to nonthermodynamic effects like variation in impact parameter. The contribution of such effects to the fluctuations have now been estimated ; the combined thermodynamic and nonthermodynamic fluctuations are in satisfactory agreement with the data . We use our analysis to argue that NA49 data are consistent with the hypothesis that almost all the observed event-by-event fluctuation in mean $`p_T`$, an intensive quantity, is thermodynamic in origin. This bodes well for the detectability of systematic changes in thermodynamic fluctuations near $`E`$. As one example, consider the ratio of the width of the true event-by-event distribution of the mean $`p_T`$ to the width of the distribution in a sample of mixed events . We called this ratio $`\sqrt{F}`$. NA49 has measured $`\sqrt{F}=1.002\pm 0.002`$ , which is consistent with expectations for noncritical thermodynamic fluctuations.<sup>3</sup><sup>3</sup>3In an infinite system made of classical particles which is in thermal equilibrium, $`\sqrt{F}=1`$. Bose effects increase $`\sqrt{F}`$ by $`12\%`$ ; an anticorrelation introduced by energy conservation in a finite system — when one mode fluctuates up it is more likely for others to fluctuate down — decreases $`\sqrt{F}`$ by $`12\%`$ ; two-track resolution also decreases $`\sqrt{F}`$ by $`12\%`$ . The contributions due to correlations introduced by resonance decays and due to fluctuations in the flow velocity are each significantly smaller than $`1\%`$ . We argue that critical fluctuations can increase $`\sqrt{F}`$ by $`1020\%`$, fifty times the statistical error in the present measurement. We have focussed on this observable because data on it has been analyzed and presented by NA49. There are, however, other observables which are much more sensitive to critical effects. For example, a $`\sqrt{F_{\mathrm{soft}}}`$, defined using only the softest $`10\%`$ of the pions in each event, may easily be affected at the factor of two level. We have estimated the magnitude of the effect of critical fluctuations on many other observables , including the event-by-event fluctuations in $`N`$, and the correlation between fluctuations in $`p_T`$ and $`N`$. We also note one observable, the multiplicity of soft pions, which may be used to detect the critical fluctuations without an event-by-event analysis. The post-freezeout decay of sigmas, which are copious and light at freezeout near $`E`$ and which decay subsequently when their mass increases above twice the pion mass, should result in a population of pions with $`p_Tm_\pi /2`$ which appears only for freezeout near the critical point . The variety of observables which should all show nonmonotonic behavior near the critical point is sufficiently great that if it were to turn out that $`\mu _E<200`$ MeV, making $`E`$ inaccessible to the SPS, all four RHIC experiments could play a role in the study of the critical point. NA49 data demonstrates very clearly that SPS collisions at $`\sqrt{s}=17`$ GeV do not freeze out near the critical point. $`E`$ has not yet been discovered. The nonmonotonic appearance and then disappearance (as $`\sqrt{s}`$ is varied) of any one of the signatures of the critical fluctuations we have described would be strong evidence for critical fluctuations. If nonmonotonic variation is seen in several of these observables, with the maxima in all signatures occurring at the same value of $`\sqrt{s}`$, this would turn strong evidence into an unambiguous discovery of the critical point. The quality of the present NA49 data, and the confidence with which we can use it to learn that collisions at $`\sqrt{s}=17`$ GeV do not freeze out near the critical point make us confident that the program of experimentation which we have described can realistically be expected to teach us much about the phase diagram of QCD, and could result in a discovery of perhaps the most fundamental feature of the landscape. Once the critical point $`E`$ is discovered, it will be prominent on the map of the phase diagram which will appear in any future textbook on QCD. ## 3 Color Superconductivity and Color-Flavor Locking As in my talk in Torino, I devote the final Section of this review to recent developments in our understanding of the low temperature, high density regions of the QCD phase diagram. In this regime, the relevant degrees of freedom are those which involve quarks with momenta near the Fermi surface. At high density, when the Fermi momentum is large, the QCD gauge coupling $`g(\mu )`$ is small. However, because of the infinite degeneracy among pairs of quarks with equal and opposite momenta at the Fermi surface, even an arbitrarily weak attraction between quarks which allows a pair of quarks with momenta $`\pm \stackrel{}{p}`$ to scatter into a pair with momenta $`\pm \stackrel{}{q}`$ renders the Fermi surface unstable to the formation of a condensate of quark Cooper pairs. Pairs of quarks cannot be color singlets, and in QCD with two flavors of massless quarks the Cooper pairs form in the color $`\overline{\mathrm{𝟑}}`$ channel . The resulting condensate gives gaps to quarks with two of three colors and breaks $`SU(3)_{\mathrm{color}}`$ to an $`SU(2)_{\mathrm{color}}`$ subgroup, giving mass to five of the gluons by the Anderson-Higgs mechanism. In QCD with two flavors, the Cooper pairs are $`uddu`$ flavor singlets and the global flavor symmetry $`SU(2)_L\times SU(2)_R`$ is intact. There is also an unbroken global symmetry which plays the role of $`U(1)_B`$. Thus, no global symmetries are broken in this 2SC phase. There must therefore be a phase transition between the 2SC and hadronic phases on the horizontal axis in Figure 1, at which chiral symmetry is restored. This phase transition is first order as is to be expected as it is characterized by a competition between chiral condensation and diquark condensation . There need be no transition between the 2SC and quark-gluon plasma phases in Figure 1 because neither phase breaks any global symmetries. However, this transition, which is second order in mean field theory, is likely first order in QCD . In QCD with three flavors of massless quarks, the Cooper pairs cannot be flavor singlets, and both color and flavor symmetries are necessarily broken. The symmetries of the phase which results have been analyzed in . The attractive channel favored by one-gluon exchange exhibits “color-flavor locking.” A condensate involving left-handed quarks alone locks $`SU(3)_L`$ flavor rotations to $`SU(3)_{\mathrm{color}}`$, in the sense that the condensate is not symmetric under either alone, but is symmetric under the simultaneous $`SU(3)_{L+\mathrm{color}}`$ rotations.<sup>4</sup><sup>4</sup>4It turns out that condensation in the color $`\overline{\mathrm{𝟑}}`$ channel induces a condensate in the color $`\mathrm{𝟔}`$ channel because this breaks no further symmetries . The resulting condensates can be written in terms of $`\kappa _1`$ and $`\kappa _2`$ where $`q_{La}^\alpha q_{Lb}^\beta \kappa _1\delta _a^\alpha \delta _b^\beta +\kappa _2\delta _b^\alpha \delta _a^\beta `$. Here, $`\alpha `$ and $`\beta `$ are color indices running from $`1`$ to $`3`$, $`a`$ and $`b`$ are flavor indices running from $`1`$ to $`3`$, and the Kronecker $`\delta `$’s lock color and flavor rotations. A condensate involving right-handed quarks alone locks $`SU(3)_R`$ flavor rotations to $`SU(3)_{\mathrm{color}}`$. Because color is vectorial, the result is to lock $`SU(3)_L`$ to $`SU(3)_R`$, breaking chiral symmetry.<sup>5</sup><sup>5</sup>5Once chiral symmetry is broken by color-flavor locking, there is no symmetry argument precluding the existence of an ordinary chiral condensate. Indeed, instanton effects do induce a nonzero $`\overline{q}q`$ , but this is a small effect . Thus, in quark matter with three massless quarks, the $`SU(3)_{\mathrm{color}}\times SU(3)_L\times SU(3)_R\times U(1)_B`$ symmetry is broken down to the global diagonal $`SU(3)_{\mathrm{color}+L+R}`$ group. All nine quarks have a gap. All eight gluons get a mass. There are nine massless Nambu-Goldstone bosons. There is an unbroken gauged $`U(1)`$ symmetry which plays the role of electromagnetism. (Under this symmetry, all the quarks, all the massive vector bosons, and all the Nambu-Goldstone bosons have integer charges.) The CFL phase therefore has the same symmetries as baryonic matter with a condensate of Cooper pairs of baryons . Furthermore, many non-universal features of these two phases correspond . This raises the possibility that quark matter and baryonic matter may be continuously connected , as shown in Figure 4. The physics of the CFL phase has been the focus of much recent work . Nature chooses two light quarks and one middle-weight strange quark, rather than three degenerate quarks as in Figure 4. A nonzero $`m_s`$ weakens those condensates which involve pairing between light and strange quarks. The CFL phase requires nonzero $`us`$ and $`ds`$ condensates; because these condensates pair quarks with differing Fermi momenta they can only exist if they are larger than of order $`m_s^2/2\mu `$, the difference between the $`u`$ and $`s`$ Fermi momenta in the absence of pairing. If one imagines increasing $`m_s`$ at fixed $`\mu `$, one finds a first order unlocking transition : for larger $`m_s`$ only $`u`$ and $`d`$ quarks pair and the 2SC phase is obtained. Conversely, as $`m_s`$ is reduced in going from Figure 2 to 3 to 4, the region occupied by the CFL phase expands to encompass smaller and smaller $`\mu `$ . For any $`m_s\mathrm{}`$, the CFL phase is the ground state at arbitrarily high density . For larger values of $`m_s`$, there is a 2SC interlude on the horizontal axis, in which chiral symmetry is restored, before the CFL phase breaks it again at high densities. For smaller values of $`m_s`$, the possibility of quark-hadron continuity as shown in Figure 4 arises. It should be noted that when the strange and light quarks are not degenerate, the CFL phase may be continuous with a baryonic phase in which the densities of all the nucleons and hyperons are comparable; there are, however, phase transitions between this hypernuclear phase and ordinary nuclear matter . The Nambu-Goldstone bosons in the CFL phase are Fermi surface excitations in which the orientation of the left-handed and right-handed diquark condensates oscillate out of phase in flavor space. The effective field theory describing these oscillations has recently been constructed . This effective theory will be useful for many purposes, including for answering the question posed in Ref. : what are the properties of the solitons, if any, in the CFL phase? The dispersion relations describing the quasiparticle excitations in the CFL phase have also received attention . One interesting possibility is that at the lowest densities at which the CFL phase exists, just above the first order unlocking transition between CFL and 2SC, the CFL phase may be a superconductor in the sense that all eight gluons are given a mass by the Meissner effect, but there may nevertheless be gapless quasiparticle excitations . Much effort has gone into estimating the magnitude of the condensates in the 2SC and CFL phases . It would be ideal if this task were within the scope of lattice gauge theory as is, for example, the calculation of the critical temperature on the vertical axis of the phase diagram. Unfortunately, lattice methods relying on importance sampling have to this point been rendered exponentially impractical at nonzero baryon density by the complex action at nonzero $`\mu `$.<sup>6</sup><sup>6</sup>6Note that quark pairing can be studied on the lattice in some models with four-fermion interactions and in two-color QCD . The $`N_c=2`$ case has also been studied analytically in Refs. ; pairing in this theory is simpler to analyze because quark Cooper pairs are color singlets. The $`N_c\mathrm{}`$ limit of QCD is often one in which hard problems become tractable. However, the ground state of $`N_c=\mathrm{}`$ QCD is not a color superconductor . This is of no concern at $`N_c=3`$, however: Shuster and Son have shown that color superconductivity persists up to $`N_c`$’s of order tens of thousands before being supplanted by the phase described in Ref. . In the absence of lattice methods, the magnitude of the gaps in quark matter at large but accessible density has been estimated using two broad strategies. The first class of estimates are done within the context of models whose parameters are chosen to give reasonable vacuum physics. Examples include analyses in which the interaction between quarks is replaced simply by four-fermion interactions with the quantum numbers of the instanton interaction or of one-gluon exchange and more sophisticated analyses done using the instanton liquid model . Renormalization group methods have also been used to explore the space of all possible effective four-fermion interactions . These methods yield results which are in qualitative agreement: the favored condensates are as described above; the gaps range between several tens of MeV up to as large as about $`100`$ MeV; the associated critical temperatures (above which the diquark condensates vanish) can be as large as about $`T_c50`$ MeV. This agreement between different models reflects the fact that what matters most is simply the strength of the attraction between quarks in the color $`\overline{\mathrm{𝟑}}`$ channel, and by fixing the parameters of the model interaction to fit, say, the magnitude of the vacuum chiral condensate, one ends up with attractions of similar strengths in different models. The second strategy for estimating gaps and critical temperatures is to use $`\mu =\mathrm{}`$ physics as a guide. At asymptotically large $`\mu `$, models with short-range interactions are bound to fail because the dominant interaction is due to the long-range magnetic interaction coming from single-gluon exchange . The collinear infrared divergence in small angle scattering via one-gluon exchange (which is regulated by dynamical screening ) results in a gap which is parametrically larger at $`\mu \mathrm{}`$ than it would be for any point-like four-fermion interaction. At $`\mu \mathrm{}`$, where $`g(\mu )0`$, the gap takes the form $`\mathrm{\Delta }256\pi ^4\mu g(\mu )^5\mathrm{exp}[3\pi ^2/\sqrt{2}g(\mu )]`$; the $`g(\mu )`$ dependence was discovered in Ref. and has now been confirmed using a variety of methods . A consequence of this result is that the magnitude of the condensates increases slowly as $`\mu \mathrm{}`$; this means that the CFL phase is favored over the 2SC phase for $`\mu \mathrm{}`$ for any $`m_s\mathrm{}`$ . The $`g`$-independent prefactor has been estimated in Refs. . As conjectured in , this prefactor is such that if this asymptotic expression is applied to accessible densities, say $`\mu 5001000`$ MeV, it predicts gaps as large as about $`100`$ MeV and critical temperatures as large as about $`50`$ MeV . The weak-coupling calculation of the gap in the CFL phase is the first step toward the weak-coupling calculation of other properties of this phase, in which chiral symmetry is broken and the spectrum of excitations is as in a confined phase, like for example the coefficients in the effective field theory of Ref. which describes the physics of the Nambu-Goldstone bosons. It is satisfying that two very different approaches, one using zero density phenomenology to normalize models, the other using weak-coupling methods valid at asymptotically high density, yield predictions for the gaps and critical temperatures at accessible densities which are in good agreement. $`T_c50`$ MeV is much larger relative to the Fermi momentum (say $`\mu 500`$ MeV) than in low temperature superconductivity in metals. This reflects the fact that color superconductivity uses attraction due to the primary, strong, interaction in the theory, rather than having to rely on much weaker secondary interactions, as in phonon mediated superconductivity in metals. Quark matter is a high-$`T_c`$ superconductor by any reasonable definition. It is unfortunate that its $`T_c`$ is nevertheless low enough that it is unlikely the phenomenon can be realized in heavy ion collisions. Neutron stars have $`T0`$, relative to the scales of Figures 1-4. If they have quark matter cores, those cores are color superconductors, in either the 2SC or CFL phase. The higher temperature regions of the phase diagram are being mapped in heavy ion collisions; we need to learn how to use neutron star phenomenology to map the high density regime. The qualitative questions to which analysis of neutron star phenomenology may yet provide answers include: Does baryonic matter change continuously into quark matter in the CFL phase as a function of increasing density (increasing depth) within a neutron star? This would require that the high density region of the phase diagram be similar to Figure 4. Or, as in Figure 3, do neutron stars have a 2SC core, perhaps with a CFL inner core? Or, of course, do neutron stars not have quark matter cores at all? Further work is required. The rate of cooling of neutron stars by neutrino emission from their cores may yield useful information: both the emission rates and the heat capacity are affected by diquark condensates and the cooling rates resulting in the 2SC and CFL phases differ . The $`r`$-mode instability is sensitive to the presence of quark matter within neutron stars ; it remains to be seen whether it differentiates between the 2SC and CFL phases. Pulsars are characterized by strong magnetic fields; further work is required to determine how the field configuration would be affected by the presence of a 2SC or CFL core , and whether the time evolution of pulsar magnetic fields would be affected. The answer to the question of whether the QCD phase diagram does or does not feature a 2SC interlude on the horizontal axis, separating the CFL and baryonic phases in both of which chiral symmetry is broken, depends on whether the strange quark is effectively heavy or effectively light. This is the central outstanding question about how to draw the high density region of the map. The central question at higher temperatures, namely where does nature locate the critical point $`E`$, also depends on the strange quark mass. Both questions are hard to answer theoretically with any confidence. The higher temperature region of the map is in much better shape, however, because the program of experimentation described in the previous Section allows heavy ion collision experiments to search for the critical point $`E`$. Theorists have described how to use phenomena characteristic of freezeout in its vicinity to discover $`E`$; this enables experimentalists to answer the question of its location convincingly. The discovery of $`E`$ would allow us to draw the higher temperature regions of the map of the QCD phase diagram in ink. Theorists have much work to do, however, before a program of astrophysical observations which would allow the inking in of the boundaries of the 2SC and CFL phases can be proposed.
no-problem/9908/astro-ph9908128.html
ar5iv
text
# The ASCA Medium Sensitivity Survey (the GIS Catalog Project): Source Counts and Evidence for Emerging Population of Hard Sources ## 1 Introduction X-ray surveys are the most direct approach to reveal the nature of sources that contribute to the Cosmic X-ray Background (CXB or XRB; see Fabian and Barcons 1989 for review). The ROSAT satellite resolved 70–80% of the CXB in the 0.5–2 keV band into discrete sources, whose majority are Active Galactic Nucleis (AGNs) (e.g., Hasinger et al. 1998; Schmidt et al. 1998). Because of the technical difficulties, imaging sky surveys in the hard X-ray band (above 2 keV), where the bulk of the CXB emission arises, were not available until the launch of ASCA (Tanaka, Inoue, and Holt, 1993). Deep/medium surveys performed with ASCA (Ogasaka et al. 1998; Ueda et al. 1998; 1999) and BeppoSAX (Fiore et al. 1999) have resolved a significant fraction (25–35%) of the CXB above 2 keV. The results indicate emergence of faint, hard X-ray sources that could be responsible in producing the CXB spectrum, which is harder than that of nearby type-I AGNs. These surveys, though novel, are limited in sky coverage. As a result, the sample size of detected sources is not sufficient to obtain a self-consistent picture about the evolution of the sources over the wide fluxes, from $`10^{11}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV) which is the sensitivity limit of HEAO1 A2 (Piccinotti et al. 1982), down to $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV), that of ASCA (e.g., Ueda et al. 1999). In particular, the broad band properties of sources at these fluxes are somewhat puzzling according to previous studies. The source counts in the soft band (0.3–3.5 keV) obtained by Einstein Extended Medium Sensitivity Survey (EMSS; Gioia et al. 1990) is about 2 times smaller than that in the hard band (2–10 keV) obtained by the Ginga fluctuation analysis (Hayashida, Inoue, & Kii 1990; Butcher et al. 1997) when we assume a power-law photon index of 1.7. This may imply presence of many hard sources at the flux level of $`10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV). Such evidence is not seen, however, in the spectrum of the fluctuation observed by Ginga, which shows a photon index of $`1.8\pm 0.1`$ in the 2–10 keV range (Butcher et al. 1997). To complement these shortcomings of deep surveys, we have been working on the project called the “ASCA Medium Sensitivity Survey (AMSS)”, or the GIS catalog project (Ishisaki et al. 1995; Ueda et al. 1997; Takahashi et al. 1998). In the project, we utilize the GIS data from the fields that have become publicly available to search for serendipitous sources. The large field of view and the low-background characteristics make the GIS instrument ideal for this purpose (Ohashi et al. 1996; Makishima et al. 1996). In this letter, we present the first results obtained from the latest version of the GIS catalog, in which the data taken from May 1993 through December 1996 are analyzed. The source list and detailed description of the catalog will be presented in a separate paper. From the serendipitous fields amounting to 106 deg<sup>2</sup>, we derived the Log $`N`$ \- Log $`S`$ relations in the 0.7–7 keV and 2–10 keV bands with the best statistical accuracy obtained so far, over the wide flux range of more than 2 decades. The AMSS sample, currently the largest sample in the 0.7–10 keV range, leads us to the best understanding of the statistical properties of sources that produce about 30% of the CXB. ## 2 Results ### 2.1 Field Selection and Data Used in the Analysis The first ASCA GIS source catalog (Ueda et al. 1999, in preparation) contains 1345 sources, including “target sources” at which the observation is aimed, detected from 369 fields observed from 1993 May to 1996 December. The analysis method for source surveys is basically the same as applied for the ASCA LSS (Ueda et al. 1998; 1999) except that the SIS data are not used in the AMSS. In the procedure we took into account the complicated responses of the XRT and the detectors. The analysis procedure consists of two steps: I. source detection and II. flux calculation. In step I, raw images are cross-correlated (smoothed) with the position dependent PSF of the XRT and the GIS and source candidates are searched in the smoothed image. In step II, we perform a 2-dimensional maximum-likelihood fitting to the raw image with a model that consists of the background and source peaks found in step I, including that of a target source. Multiple observations of the same or overlapped fields are combined in the analysis. The following selection criteria are applied for the catalog: (1) the Galactic latitude $`|b|`$ is higher than $`10^{}`$, (2) the time-averaged count rate is less than $`0.8`$ c s<sup>-1</sup> per sensor, (3) the exposure is longer than 5000 sec, and (4) the 2-dimensional fit in step II is successful (reduced $`\chi ^2<1.7`$). We also excluded any fields for which we could not model the surface profile adequately, such as the fields of bright cluster of galaxies. In this paper, to limit our interest on the study of extra-galactic sources, we further selected the fields of $`|b|>20^{}`$, excluding observations of nearby galaxies of large angular size, star forming regions, and supernova remnants. Finally, we discarded the target sources from the source list to construct a complete sample consisting of only serendipitous sources. The sample contains 714 sources (above 5 $`\sigma `$ detection), of which 696, 323, and 438 sources are detected in the 0.7–7 keV (total), 2–10 keV (hard), and 0.7–2 keV (soft) band, respectively. The number of sources detected both in the total and hard bands, total and soft bands, and hard and soft bands are 320, 423, and 266, respectively, and 266 sources are detected in all the survey bands. The total sky area covered amounts to 106 deg<sup>-2</sup> and the sensitivity limits are $`5\times 10^{14}`$, $`7\times 10^{14}`$, and $`2.6\times 10^{14}`$ erg s<sup>-1</sup> cm<sup>-2</sup> for the 0.7–7 keV, 2–10 keV, and 0.7–2 keV survey band, respectively. For each field, we corrected the count rates for the Galactic absorption, whose column density is estimated from H I observations (Dickey & Lockman 1990). The LSS sources (Ueda et al. 1999) are not included in the sample, whereas most of the GIS data of the deep surveys are included there. ### 2.2 The Source Spectra Figure 1 shows the correlation between the 0.7–7 keV flux and the hardness ratio in the 0.7–10 keV range, $`HR1`$, defined as $`(HS)/(H+S)`$, where $`H`$ and $`S`$ represent the count rate within radius of $`5^{}.9`$ around the source (corrected for vignetting, see Ueda et al. 1999) in the 2–10 keV and 0.7–2 keV bands, respectively. Due to the limited photon statistics, however, these spectral information of individual sources is subject to a large statistical error, typically by 0.03, 0.1, and 0.2 (1$`\sigma `$) at fluxes of $`10^{11}`$, $`10^{12}`$, and $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup>, respectively. Hence, to study the average properties of the source spectra, we calculated the average value of $`HR1`$ (weighting with their errors) in several flux ranges from sources detected in the total band. The results are plotted in Figure 1 with crosses. Since the sensitivity is given in count rate rather than in flux, we sort sources by count rate (not by flux), to avoid to introduce any selection effects that very hard (or soft) sources are difficult to detect at the flux close to the sensitivity limit of each observation. The dashed curves in Figure 1 show conditions that give the same count rate, and sources located in the regions between the two curves are used for calculation of average hardness ratio. It is clear from Figure 1 that the average spectrum becomes harder with a decreasing flux. The corresponding photon index (assuming a power law over the 0.7–10 keV band with no absorption) changes from 2.1 at the flux of $`10^{11}`$ erg s<sup>-1</sup> cm<sup>-2</sup> to 1.6 at $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–7 keV). ### 2.3 Log $`N`$ \- Log $`S`$ Relation Using the sample, we derived the Log $`N`$ \- Log $`S`$ relation independently in the three survey bands, 0.7–7 keV, 2–10 keV, and 0.7–2 keV, in the following manner. For each field, we calculated the observed area as a function of count rate, $`\mathrm{\Omega }(S)`$, following the same procedure described in Ueda et al. (1999). In this process, we estimate the significance of detection expected for the given flux at every positions, utilizing the fitting model used in step II of the source finding procedure and taking into account the presence of the target source. In calculation of $`\mathrm{\Omega }(S)`$ we excluded the regions where the spill of the PSF from the target source is higher than 50% of the background, and did not count any sources detected within the excluded region. Then, after summing up from all the selected fields $`\mathrm{\Omega }(S)`$, and $`N(S)`$, the number of sources in flux bin of $`dS`$, we calculated the “observed” Log $`N`$ \- Log $`S`$ relation in the differential form by dividing $`N(S)`$ by $`\mathrm{\Omega }(S)`$. To evaluate possible systematic errors in the observed Log $`N`$ \- Log $`S`$ relations, we next performed a large number of Monte Carlo simulations with roughly the same exposure distribution as the real data: we created 1400 simulated image data, consisting of 400 pointings with an exposure of 20 ksec, 300 with 30 ksec, 500 with 40 ksec, 50 with 70 ksec, and 50 with 90 ksec. The Log $`N`$ \- Log $`S`$ relation input to the simulations is taken to be the same as the one we derived from the AMSS. We then applied the same analysis procedure for these simulated data as for the real data, and compared the derived parameters with the input parameters. We found an excess of the output source counts relative to the input ones, which is about 10% (20%) at $`S=2\times 10^{13}`$ ($`8\times 10^{14}`$) erg s<sup>-1</sup> cm<sup>-2</sup> and 20% (40%) at $`S=7\times 10^{14}`$ ($`3\times 10^{14}`$) erg s<sup>-1</sup> cm<sup>-2</sup> in the total (soft) band, while such deviation is negligible in the hard band. The deviation and its energy dependence are also seen in the case of the LSS (Ueda et al. 1999) and can be explained by the effect of the source confusion, because the positional resolution of the GIS becomes worse toward lower energies. Thus, we corrected the observed differential Log $`N`$ \- Log $`S`$ relation for this effect with a multiplicative factor $`f`$, derived by the simulation, which has a form of $`f=1/[\mathrm{\Delta }\times (\mathrm{log}S\mathrm{log}S_0)/(\mathrm{log}S_1\mathrm{log}S_0)+1]`$ for $`S<S_0`$, and f = 1 for $`S>S_0`$, where $`\mathrm{\Delta }`$ = 0.2 (0.4), $`S_0=5.9\times 10^{13}(2.2\times 10^{13})`$, and $`S_1=7.4\times 10^{14}(2.7\times 10^{14})`$, for the total (soft) band survey, respectively. These simulated data are also used to evaluate the systematic error due to the source confusion in determination of $`HR1`$: the amount of the bias is estimated to be $`0.03`$ at $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–7 keV) and is negligibly small in brighter flux level. This bias has been taken into account in deriving the average hardness ratio in Figure 1. Since we impose the maximum count rate in the field-selection criteria, very bright sources are intentionally excluded in our sample: this means there is an upper flux limit above which the sample becomes incomplete. The maximum count rate, 0.8 c/s/sensor including the background and the target sources, indicates the upper flux limit is higher than $`(11.5)\times 10^{11}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV) for most of the fields, although it depends both on the spectrum and the brightness of the target source. In integrating the differential Log $`N`$ \- Log $`S`$ relations, we set the upper flux limits at $`10^{11}`$ erg s<sup>-1</sup> cm<sup>-2</sup> for the 0.7–7 keV and 2–10 keV surveys, and $`5\times 10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> for the 0.7–2 keV survey, and used previous results by HEAO1 A2 (Piccinotti et al. 1982) as integral constants at the bright flux ends assuming a photon index of 1.7. The uncertainty in the completeness close to these limits does not affect our discussion below, which are made mainly based on the fluxes fainter than $`10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–7 keV and 2–10 keV). The large sample size enables us to derive Log $`N`$ \- Log $`S`$ relations separately for sources with different spectra. In practice, considering the large statistical error in the hardness ratio for an individual source, we divide the sample into two, the “soft source sample”, consisting of sources with $`HR1`$ smaller than –0.028, which corresponds to a photon index larger than 1.7, and the “hard source sample”, with a photon index smaller than 1.7. In the conversion from count rate into flux, we assumed a photon index of 1.6 (1.6) for the hard source sample and 1.9 (1.6) for the soft source sample in the 0.7–7 keV (2–10 keV) survey. These photon indices come from the average spectrum of each sample after correcting for biases of purely statistical origin, based on the simulation. Figure 2 shows the integral Log $`N`$ \- Log $`S`$ relations in the 0.7–7 keV survey band for the soft source sample (red curve), the hard source sample (blue curve), and the sum (black curve). The figure clearly demonstrates that sources with hard energy spectra in the 0.7–10 keV range are rapidly increasing with decreasing fluxes, compared with softer sources. The ratio of the hard source sample (with a photon index of less than 1.7) to the soft source sample changes from about 20% at the flux of $`10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–7 keV) to about 50% at $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup>. We confirmed by the simulations that any statistical biases does not produce such dramatic change as a function of flux. For the 2–10 keV survey, we show the sum of the soft and hard source samples in Figure 3 (black curve). ## 3 Discussion First, for confirmation, we compared the Log $`N`$ \- Log $`S`$ relation in the 0.7–2 keV band with the results by ROSAT in almost the same band (0.5–2 keV) (Hasinger 1998 and references therein). We found a good agreement within the statistical error at $`S>1\times 10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–2 keV) between the two, whereas we see a slight excess of the ASCA source counts by about 10–20% at $`S=3\times 10^{14}1\times 10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–2 keV), which was also reported from the LSS (Ueda et al. 1999). Using the AMSS data, we found that these excess can be explained by contribution of hard sources that have an (apparent) photon index less than 1.7 in the 0.7–10 keV range, whose number density increases more rapidly than that of softer sources toward fainter fluxes in the 0.7–2 keV band. Thus, as is discussed in Ueda et al. (1999), it is plausible that the difference between the ASCA and ROSAT results can be accounted for by the presence of hard (or absorbed) sources, which are more easily detectable by ASCA than by ROSAT, due to their difference of energy dependence of the effective area even within the similar energy band. The present results of the Log $`N`$ \- Log $`S`$ relations in the three survey bands are all consistent with the LSS results (Ueda et al. 1999) within the statistical errors. The 2–10 keV source counts is also consistent with that by Cagnoni, Della Ceca, & Maccacaro (1998), who utilized the GIS2 data from about one fifth of the data set analyzed here, and with the fluctuation analysis of the ASCA SIS data by Gendreau, Barcons & Fabian (1998). In Figure 3, we show the constraints from the fluctuation analysis by Ginga (Butcher et al. 1997). As noticed, below the flux of $`10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV), our direct source counts gives somewhat lower values than Ginga, although it is barely consistent within the 90% statistical errors. The increasing fraction of hard sources toward fainter fluxes seen in Figure 2 accounts for the evolution of the average source spectra shown in Figure 1. The spectral evolution is also consistent with the comparison of the source counts between the hard and the soft band. In Figure 3, we compare the source counts in the hard band with that in the soft band, both obtained from the AMSS. In the figure, the 0.7–2 keV fluxes are converted into the 2–10 keV fluxes assuming the two photon indices (1.6 and 1.9). As is clearly seen, the source counts in the hard band survey increases more rapidly than that in the soft band survey toward the sensitivity limit. As a result, the hard band source counts at $`S10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV) matches the soft band one when we assume a photon index of 1.6, whereas at brighter level of $`S=4\times 10^{13}10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV), we have to use a photon index of about 1.9 to make them match. These indices, which should represent average source spectra at each flux level, well coincide with those directly derived from Figure 1. It is also interesting that the slope of the Log $`N`$ \- Log $`S`$ relation in the hard- and soft-source sample in the total band survey (Figure 2) is similar to that of the Log $`N`$ \- Log $`S`$ relation in the hard- and soft-band survey (Figure 3), respectively, at the flux range of one decade above the sensitivity limit. This is expected if we roughly consider that the sources in the hard (soft) source sample correspond to those detected in the hard (soft) survey. As discussed above, all the present results from the AMSS are perfectly consistent with one another. Consequently, the derived Log $`N`$ \- Log $`S`$ relations covering the 0.7–10 keV range, determined by direct source counts, has now solved the puzzle of discrepancy of the source counts between the soft (EMSS) and the hard band (Ginga and HEAO1). They are now reconciled by the two facts: (1) the source counts in the hard band obtained by the AMSS is smaller than the best-fit value of the Ginga results at $`S=4\times 10^{13}1\times 10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV) and (2) that in the soft band, which is consistent with the recent ROSAT surveys such as RIXOS (Mason et al. 1999), gives larger source counts by $``$30% than the EMSS results that include Galactic objects. As seen from Figure 1, the average spectrum of sources at $`S=3\times 10^{13}3\times 10^{12}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (0.7–7 keV) has a photon index of 1.8–2.1 in the 0.7–10 keV range, indicating that the contribution of hard sources (such as heavily absorbed AGNs) is not significant yet at this flux level to reproduce the CXB spectrum. This fact can be connected with the “soft” spectrum of the fluctuation observed with Ginga, which shows a photon index of 1.8$`\pm 0.1`$ in the 2–10 keV range (Butcher et al. 1997). The emerging population of hard sources seen in Figure 2 can be understood by increasing contribution of absorbed AGNs, which should become more significant toward fainter flux level due to the K-correction effect as is discussed in the AGN synthesis model (Awaki et al. 1991; Comastri et al. 1995). In fact, optical identification of the ASCA LSS shows that most of the hard (or absorbed) sources are narrow-line or weak broad-line AGNs, and that their contribution is comparable to that of unabsorbed AGNs at the 2–10 keV flux of $`2\times 10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (Akiyama et al. 1999). Since the number of sources in the LSS sample is limited, the optical identification of the AMSS sample is crucial to reveal the evolution of each population, particularly that of absorbed AGNs, which are difficult to detect in the soft band. The integrated spectrum of sources with fluxes above $`10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV) is not hard enough to completely account for the CXB spectrum. Our results thus predict that further hardening is necessary at fainter flux levels. This can be confirmed by future missions such as XMM and Chandra. On the other hand, at the brighter flux range, $`10^{13}10^{11}`$ erg s<sup>-1</sup> cm<sup>-2</sup> (2–10 keV) where we need to cover large area to overcome the small surface number densities, the AMSS provides the best opportunity for statistical studies of X-ray sources in the universe. We are grateful to Prof. H. Inoue for stimulating discussion. We thank members of the ASCA team for their support in satellite operation and data acquisition, and Dr. H. Kubo for the help in analysis.
no-problem/9908/cond-mat9908395.html
ar5iv
text
# Charging effects in a quantum wire with leads ## I Introduction Electron transport in quantum wires (QWs) now attracts considerable interest because of the fundamental importance of the electron-electron (e-e) interaction in one-dimensional (1D) systems. In addition, it is expected that the e-e interaction may produce important effects in the transport that would be attractive for applications. However, up to now there is no clear and unambiguous knowlege of which effects of e-e interaction are observable under the realistic conditions of a QW with leads and what the reasons are for the deviations from conductance quantization (observed experimentally ), and very little is known about the electron transport under far from equilibrium conditions. In order to understand the transport properties, it is necessary to know the spatial distribution of the electric potential and the electron density in the structures under investigation. Many arguments and facts show that the leads play an essential role in the conductance, if the transport is investigated by measuring the electric current in an external circuit and by measuring the voltage drop between the leads. It is obvious that the interaction of the QW with the leads is not weak. In the present paper we investigate this interaction, considering the QW and the leads as a unified system. An important problem that arises is how the electric field and the electron density are distributed at equilibrium (i.e., in the ground state), and how they are redistributed when an external voltage is applied. We show that even in the equilibrium state the QW acquires a charge and a contact potential difference between the QW and the leads. This phenomenon is similar to the well known contact potential difference in classical conductor systems. The essential difference between the low-dimensional system under consideration and classical conventional 3D conductors is that the contact potential is not screened over a finite length but spreads over a length determined by the geometrical size of the structure. The contact potential difference is determined by the difference in the chemical potentials $`\mathrm{\Delta }\mu `$ of electrons in the QW and in the leads, when these subsystems are considered independently. By analyzing the chemical potentials we find that three cases are possible, depending on the wire radius and the background charge density: (i) the QW is charged positively, (ii) the QW acquires a negative charge, or (iii) the QW remains neutral as a whole. We develop a model of a QW with leads in which the lead-wire interaction is taken into account. It is based on a Hartree-Fock approach for the electrons in the QW and the representation of leads (which are considered as electron reservoirs) in a way that takes into account their 3D nature, but requires only a 1D calculation. Using this model we investigate the distribution of the electron density and the electric potential at equilibrium, as well as under far from equilibrium conditions. In particular, we emphasize the role of the exchange interaction effect on these quantities. If in the absence of coupling the chemical potential of the wire exceeds that of the reservoir, $`\mathrm{\Delta }\mu >0`$, the coupled system expels electrons from the QW into the leads, when equilibrium is established. As a result a potential well appears in the QW. The wire acts as a charge donor to the reservoir. If $`\mathrm{\Delta }\mu <0`$, electrons are attracted by the wire, and a potential barrier arises between the QW and the leads. The wire acts in this case as an acceptor. In the case where $`\mathrm{\Delta }\mu =0`$, the electron density is redistributed only within the QW and only Friedel oscillations arise near the contacts. The Friedel oscillations appear in all cases; however, they are superimposed on a much more slowly varying potential created by the acquired charge. The slowly varying potential has an amplitude that is large compared to the Friedel oscillations and thus dominates the scattering processes. The exchange interaction strongly affects the potential shape and somewhat enhances the Friedel oscillation amplitude. When an external voltage is applied, the chemical potentials in the electron reservoirs are shifted relative to each other, disturbing the electron flows in the QW. As this takes place, the electron density, the potential, and the exchange energy are changed self-consistently. The importance of the electrostatic potential distribution in quantum wires with leads, especially for the investigation of time-dependent transport, and for nonlinear transport, has been emphasized previously, but quantitative calculations have to our knowledge not been reported thus far. Under far from equilibrium conditions, when the applied voltage exceeds the Fermi energy in the QW, the electron density is substantially redistributed between the QW and the reservoirs, giving rise to a very strong variation in the potential landscape. In turn the potential produces a variation in the electron density. The connection between the electron density and the potential is very important for the understanding of nonlinear transport. The need of a self-consistent treatment has been emphasized by Landauer. Within the scattering approach, in the weakly nonlinear regime, it has been investigated by Christen and one of the present authors and by Ma, Wang and Guo. The calculation of the present paper allows us to investigate the strongly nonlinear transport in the system under consideration. If the applied voltage is high enough, the self-consistent connection of charge and potential gives rise to an instability of the electron density distribution and ultimately leads to multistability of the electron states in the QW. This means that several stable states with different spatial distributions of the electron density and the potential are possible at a given applied voltage. The paper is organized as follows. In Sec. II the chemical potential difference between the decoupled QW and the 2D electron reservoir is analyzed. Section III describes the model of the QW with leads. Section IV contains the results of the numerical calculations of the electron density and potential distribution in the QW with leads. In Sec. VI the multistability of the electron states is described that appears for far from equilibrium conditions. ## II Contact potential difference To be specific we consider a QW connecting two regions of a 2D electron gas. We assume that there are no nearby gates and that all electric field lines emanate and terminate either on the wire or on the 2D electron gas. QW structures of this kind are produced by etching of heterostructures with a 2D electron gas. Such structures are widely used in experiments. First we investigate such a QW separately from 2D electron reservoirs. The uncoupled wire is charge neutral. The electron charge is concentrated inside the QW, while the compensating positive charge of the impurities is really located in the immediate vicinity of the QW or at its surface. The decoupled QW and the electron reservoir have their own chemical potentials $`\mu _{1\mathrm{D}}`$ and $`\mu _{2\mathrm{D}}`$ which are generally not equal each other. We are interested in the chemical potential difference $`\mathrm{\Delta }\mu `$ between the QW and the reservoir. According to Seitz’s theorem the chemical potential in the QW is determined by the Fermi energy of noninteracting electrons and the self-energy $`\mathrm{\Sigma }(k_F)`$ which takes into account the e-e interaction, $$\mu _{1\mathrm{D}}=\epsilon _0+\epsilon _F+\mathrm{\Sigma }(k_F).$$ (1) Here $`\epsilon _0\pi ^2\mathrm{}^2/(2ma^2)`$ is the first subband energy caused by transverse confinement ($`a`$ is the QW radius, $`m`$ is the effective mass of electrons). $`\mathrm{\Sigma }(k_F)`$ contains the contributions arising from the exchange and correlation interaction as well as from the electron interaction with the positive background charge. The exchange and correlation energy was investigated in the recent paper of Calmels and Gold using the self-consistent theory of Singwi, Tosi, Land and Sjölander for the case where only the lowest subband is occupied. The Hartree energy is easily estimated if we assume that the positive charge is located at the surface of the QW. These calculations lead to the following expression for the chemical potential $`\mu _{1\mathrm{D}}`$ of the QW in terms of the dimensionless parameters $`r_s=1/(2a_Bn)`$ and $`\beta =a_B/a`$ (with $`a_B`$ the effective Bohr radius and $`n`$ the 1D electron density): $$\frac{\mu _{1\mathrm{D}}}{R_y}\frac{\pi ^2}{2}\beta ^2+\frac{\pi ^2}{16r_s^2}+\frac{\mathrm{\Sigma }_{\mathrm{xc}}}{R_y}\frac{B_H}{r_s},$$ (2) where $`R_y`$ is the effective Rydberg. In Eq. (2) the first term is the lowest subband energy, the second term is the kinetic energy, and the third term represents the exchange and correlation energy. There are two expressions for the exchange-correlation term depending on whether $`r_s<1`$ or $`r_s>1`$: $$\frac{\mathrm{\Sigma }_{\mathrm{xc}}}{R_y}\beta \frac{5.57\pi 4\beta r_s}{2\pi ^2}\text{if}r_s<1$$ and $$\frac{\mathrm{\Sigma }_{\mathrm{xc}}}{R_y}\frac{1.84}{r_s}\left[\mathrm{ln}\frac{2\beta r_s}{\pi }+0.7115\right]\text{if}r_s>1.$$ The last term in Eq. (2) is the Hartree energy of the electron interaction with the positive background; $`B_H`$ is a numerical factor that depends on the radial distribution of the electron density. If the electron density is distributed uniformly, $`B_H1/3`$. The chemical potential in the 2D reservoir can easily be obtained from the known expression for the electron energy as a function of the density parameter $`R_s=(\pi a_B^2N_{2\mathrm{D}})^1`$ (where $`N_{2\mathrm{D}}`$ is the 2D electron density). In the high density case, $`R_s<\sqrt{2}`$, one obtains $$\begin{array}{cc}\hfill \frac{\mu _{2\mathrm{D}}}{R_y}=& \left(\frac{\pi a_B}{d}\right)^2+2R_s^21.80R_s^10.38\hfill \\ & 0.0863R_s\mathrm{ln}R_s+0.519R_s,\hfill \end{array}$$ (3) where $`d`$ is the thickness of the 2D layer. Using Eqs. (2) and (3), the chemical potential difference $`\mathrm{\Delta }\mu =\mu _{1\mathrm{D}}\mu _{2\mathrm{D}}`$ is calculated as a function of the QW radius and the density parameter $`r_s`$. The results of these calculations are illustrated in Fig. 1, where the $`\mathrm{\Delta }\mu `$ dependence on the wire radius $`a`$ is shown for various values of $`r_s`$. Here we consider $`r_s`$ as an independent parameter because the background charge density depends on external factors, such as the charge absorbed at the wire surface. Figure 1 shows that in sufficiently thin wires the chemical potential is higher than in the reservoir. However, with increasing radius of the QW the chemical potential in the wire can become lower than in the reservoir. The chemical potential difference $`\mathrm{\Delta }\mu `$ is caused by all energy components contributing to the chemical potentials in the QW and in the reservoir. As an example it is instructive to consider the estimations for the specific case where the QW diameter is equal to the 2D layer thickness, $`d=2a`$, and the background charge density per unit area, $`N_{2\mathrm{D}}`$, is the same in the reservoir and in the QW. The latter means that $`n=2aN_{2\mathrm{D}}`$. In this case the energies contributing to the chemical potentials are estimated as follows. The confinement energy in the QW is approximately twice that in the reservoir. Hence the confinement energy causes $`\mu _{1\mathrm{D}}`$ to rise with respect to $`\mu _{2\mathrm{D}}`$. The ratio of kinetic energy in the QW to that in the reservoir is $`\epsilon _{F1}/\epsilon _{F2}an`$. Since the QW is supposed to be a 1D system, the product $`an`$ must be small. Hence the Fermi energy $`\epsilon _{F1}`$ in the QW is noticeably smaller than the Fermi energy $`\epsilon _{F2}`$ in the reservoir. This results in lowering $`\mu _{1\mathrm{D}}`$ relative to $`\mu _{2\mathrm{D}}`$. The ratio of the exchange-correlation energies in the case of $`r_s<1`$ is estimated as $`\epsilon _{\mathrm{xc1}}/\epsilon _{\mathrm{xc2}}(an)^{1/2}`$. The exchange-correlation energy in the QW is seen to be larger than that in the reservoir. Taking into account that the exchange-correlation energy is negative, we conclude that it lowers $`\mu _{1\mathrm{D}}`$ with respect to $`\mu _{2\mathrm{D}}`$. If the wire is now coupled to the 2D reservoir, we can distinguish three cases. For $`\mathrm{\Delta }\mu >0`$, electrons are transferred from the QW to the reservoir, so that the positive background charge dominates the electron charge in the wire. If $`\mathrm{\Delta }\mu <0`$, the reservoir supplies electrons to the QW producing an excess negative charge. When $`\mathrm{\Delta }\mu =0`$, the electron density is not redistributed between the wire and reservoir. The electron density redistribution continues until an equilibrium state is attained in the whole system such that there exists a uniform electrochemical potential. As this takes place, a charge and a built-in electric field $`E_0`$ appear in the QW. If the reservoir conductivity is high, the built-in field satisfies the condition $`e_0^{\mathrm{}}𝑑xE_0(x)=\mathrm{\Delta }\mu `$. Such a charge transfer is similar to what happens when a contact potential difference appears in classical 3D systems. However, an essential difference is in the distance over which the contact field is screened. In the 3D case the contact field is screened over a finite length (e.g., the Debye length or the Thomas-Fermi screening length). In the mesoscopic structure considered here, the contact field is produced by the charge, one part of which is situated in the 1D wire and the other on the surface of the reservoir adjacent to the wire. One can easily see, that owing to the 3D nature of the electric field, any distribution of charges in a QW cannot screen the contact field over a finite distance. Thus the question arises of how the charge density and the electric field are distributed in the QW and over the reservoir surface. In the case of a high enough electron density, the interaction effects are not strong, and this problem can be solved analytically using the Thomas-Fermi approximation. In the present paper, we study this problem using a numerical solution of the Schrödinger equation coupled to the Poisson equation within the Hartree-Fock approximation. ## III Model Finding the self-consistent electron density and the potential in a QW coupled to electron reservoirs is a rather complicated problem because the electron density redistribution between the QW and the reservoirs produces a strong variation of the electron energy in the QW. The energy variation is estimated as $`\mathrm{\Delta }\mu `$, which is shown above to be of the order of the Fermi energy. In essence, the QW and the reservoirs should be considered as a unique quantum system. To our knowledge, such a problem has not been studied to date. In the present paper, we investigate it using a simplified model based on the Hartree-Fock approximation, which allows one to take adequately into account the charge accumulated in the QW both for the equilibrium state and under the far from equilibrium conditions appearing when an external voltage is applied. This approach allows one also to study the exchange interaction effect on the charge accumulated and the electric potential. However, it does not take into account the electron correlation energy. The ratio of the correlation energy to the exchange energy depends on the electron density. For 1D conductors this ratio can be estimated according to Ref. . The correlation energy is negligible when $`r_s1`$. Under this condition the Hartree-Fock approximation is justified. It is instructive to estimate numerically the number of electrons in a GaAs QW when $`r_s=0.5`$. The Fermi energy and the electron density are respectively $`\epsilon _F`$ 4 meV and $`n6\times 10^5`$ cm<sup>-1</sup>. Our computation procedure works well when the QW length is not too large in comparison with the Fermi wavelength. If one puts the QW length equal to $`0.3\mu `$m, the total number of electrons in the QW is estimated as about 18. Thus the system contains about ten electrons in a QW open to reservoirs. The number of electrons really existing in the wire and their density distribution is determined by the QW length, the background charge, the applied voltage, and the e-e interaction energy. The parameters of the QW estimated above are easily realizable in experiment. In order to investigate cases with different relative positions of the chemical potentials in the QW and reservoirs, we introduce a positive background charge density $`en_b`$ in the QW, which is considered as a parameter of our model. By varying $`en_b`$ it is possible to realize any relative position of the chemical potentials of the uncoupled system. Charges on the QW surface are not taken into account in our present consideration. Another simplification is that only the lowest subband in the QW is considered. A 1D sketch of the energy diagram of a QW with leads and the electron flows (in a far from equilibrium situation) is shown in Fig. 2. Here $`U_0`$ is the confinement energy in the QW and $`\mu _\pm `$ are the chemical potentials in the reservoirs. The positions of $`\mu _+`$ and $`\mu _{}`$ relative to the conduction band bottom of the corresponding reservoirs are fixed because the electron gas in the reservoirs is incompressible. However, $`\mu _+`$ and $`\mu _{}`$ are shifted relative to each other in the presence of an applied voltage $`V_a`$. The following energies contribute to the potential shape of the structure: the confinement energy in the QW; the potential produced by the external voltage source; the Hartree and exchange energies in the QW. In the reservoirs ($`|x|>L/2`$, $`L`$ being the wire length) the electron density is taken to be so high that the e-e interaction energy can be ignored in comparison with the kinetic energy. Moreover, the reservoirs are assumed to be ideally conducting and hence can be treated as equipotentials. Inside the QW the e-e interaction potential $`U(x,x^{})`$ is determined by both the direct interaction of electrons with each other and the indirect interaction via image charges induced by electrons in the reservoirs. Because of this, the interaction potential depends on the coordinates $`x,x^{}`$ of the interacting electrons separately, rather than simply on their difference. In the QW the one-electron wave functions $`\psi _{r,k_r,s}(x)`$ are characterized by quantum numbers: $`r=\pm 1`$ ($`r=+1`$ indicates electrons incident on the QW from the left reservoir and $`r=1`$ corresponds to electrons incident on the QW from the right reservoir); $`k_r`$ is a wave number in the left ($`r=+1`$) or right ($`r=1`$) reservoir; $`s`$ is a spin variable. In this paper we ignore changes in $`s`$ and suppose that the states with opposite spins are equally occupied. The effects of possible spin polarization in a QW will be considered elsewhere. The electron transport in the reservoirs, close to the transition between the reservoir and the QW, is modeled by a 1D Schrödinger equation without interaction. This model allows one to simulate adequately the transmission probability between the reservoirs and the QW. The value of the transmission probability calculated in this way as compared with that for a true 2D to 1D transition at the interface of the reservoir and the wire has been investigated by direct comparison of the transmission probabilities of these two geometries. The difference depends on the energy but does not exceed 15% even close to the transmission threshold. The physical reason for the success of the purely 1D model is that near the chemical potential level, of all the electron waves in the reservoir only those couple effectively to the wire that have a wave vector that is nearly parallel to the wire axis (collimation effect ). We also emphasize that in the problem under investigation the electron density of the 1D Schrödinger problem is needed only within the wire. It is unimportant for the reservoirs because they are treated as equipotentials and therefore the calculated electron density distribution in the reservoirs does not directly affect the e-e interaction in the QW. The calculated electron wave functions in the reservoirs affect only the transmission probability through the contacts. In the reservoirs, $`\psi _{r,k_r,s}(x)`$ is thus $$\psi _{r,k_r,s}(x)=\{\begin{array}{cc}\mathrm{exp}[ik_r(rx+L/2)]+R_r\mathrm{exp}[ik_r(rx+L/2)]\hfill & \text{if}rx<L/2\hfill \\ T_r\mathrm{exp}[ik_r^{}(rxL/2)]\hfill & \text{if}rx>L/2,\hfill \end{array}$$ (4) where $`(k_r^{})^2=k_r^2+r\times 2meV_a/\mathrm{}^2`$, and $`V_a`$ is the applied voltage. In the QW, $`\psi _{r,k_r,s}(x)`$ is determined by the equation $$\frac{\mathrm{}^2}{2m}\frac{d^2\psi _{r,k_r,s}}{dx^2}+\left[U_0(x)U_{\mathrm{ext}}(x)+U_H(x)+\widehat{H}_{\mathrm{ex}}\right]\psi _{r,k_r,s}=\epsilon _r(k_r)\psi _{r,k_r,s}.$$ (5) In Eq. (5) the potential energy has the following components. $`U_0(x)`$ is an effective potential that simulates the electron confinement in the QW. In the simplest case, we can assume that $`U_0(x)=U_0=\mathrm{const}`$ for $`|x|<L/2`$ and $`U_0(x)=0`$ for $`|x|>L/2`$. $`U_H`$ is the Hartree energy, $$U_H(x)=\underset{L/2}{\overset{L/2}{}}𝑑x^{}U(x,x^{})\left[n(x^{})n_b\right],$$ (6) with $`n_b`$ being the positively charged background density, $`n(x)`$ being the electron density, $$n(x)=\underset{r=\pm ,s}{}\underset{0}{\overset{\mathrm{}}{}}\frac{dk_r}{2\pi }f(k_r)|\psi _{r,k_r,s}|^2,$$ (7) and $`f(k_r)`$ the electron distribution function in the reservoirs. In Eq. (6) we assume for simplicity that the radial component of the background charge density is the same as the electron density. The e-e interaction potential $`U(x,x^{})`$ that appears in Eq. (6) depends on the spatial configuration of the leads. In what follows the numerical calculations are carried out for the case where the leads are represented as two plates perpendicular to the QW. This configuration is convenient for further calculations because in this case a relatively simple analytical expression is obtained for $`U(x,x^{})`$ This form of the interaction potential allows one to take into account not only the direct Coulomb interaction of electrons but also their interaction via image charges induced on the lead surfaces. The interaction potential is $$U(x,x^{})=\frac{e^2}{ϵL}\underset{0}{\overset{\mathrm{}}{}}\frac{dy}{\mathrm{sinh}y}|\chi _y|^2\{\begin{array}{ccc}\hfill \mathrm{sinh}\left[y\left(1/2+\xi \right)\right]\mathrm{sinh}\left[y\left(1/2\xi ^{}\right)\right]& \mathrm{if}& \xi <\xi ^{}\hfill \\ \hfill \mathrm{sinh}\left[y\left(1/2\xi \right)\right]\mathrm{sinh}\left[y\left(1/2+\xi ^{}\right)\right]& \mathrm{if}& \xi >\xi ^{},\hfill \end{array}$$ where $`\xi =x/L`$ and $`\chi _y`$ is the Fourier transform of the radial density, which is taken to be $`\chi _y=\mathrm{exp}\left[(ay/2L)^2\right]`$. Using the analytical expression for $`U(x,x^{})`$, instead of direct solution of the 3D Poisson equation, highly facilitates computations. $`\widehat{H}_{\mathrm{ex}}`$ is the exchange energy operator, $$\widehat{H}_{\mathrm{ex}}(x)\psi _{r,k_r,s}=\underset{L/2}{\overset{L/2}{}}𝑑x^{}U(x,x^{})n_{\mathrm{ex}}(x,x^{})\psi _{r,k_r,s}(x^{}),$$ where $$n_{\mathrm{ex}}(x,x^{})=\underset{r=\pm }{}\underset{0}{\overset{\mathrm{}}{}}\frac{dk_r}{2\pi }\left[\psi _{r,k_r,s}^{}(x^{})\psi _{r,k_r,s}(x)\right]f(k_r).$$ For the reservoir configuration that we consider here, the ”external potential” is a linear function of $`x`$: $`U_{\mathrm{ext}}=eV_a(x/L+1/2)`$. The energy $`\epsilon _r(k_r)`$ in Eq. (5) is expressed in terms of the wave vector $`k_r`$ and the applied voltage $`V_a`$, $$\epsilon _r(k_r)=\frac{\mathrm{}^2k_r^2}{2m}+eV_a\delta _{r,1},$$ where we assume that the energy reference is fixed at $`x=\mathrm{}`$. In addition, we require continuity of the wave functions determined by Eqs. (5) and (4) and their derivatives at the reservoir-wire interfaces $`x=\pm L/2`$. The distribution functions $`f(k_r)`$ in the reservoirs are taken in the form of the Fermi functions with the temperature $`T`$ considering the fact that the Fermi level in the right reservoir is shifted down by $`eV_a`$ with respect to the left one. Inside the QW no distribution functions are assigned. The electron distribution over the energy is determined by the electron flows from the left and right reservoirs and the interaction processes inside the QW. The external voltage produces a variation of the electron flows, as a consequence of which the electron density is changed. Ultimately, this results in the self-consistent variation of both the electron states and their occupation for both the left and right moving particles. The wave functions that we consider in this work are characterized by a continuous quantum number $`k_r`$ Hence, $`\psi `$ should be considered as a function of two variables $`x`$ and $`k_r`$. Equation (5) is an integro-differential equation with respect to the variable $`x`$ and an integral equation with respect to the variable $`k_r`$. We develop a numerical scheme for the solution of this equation on a grid spanning the two variables. The computation method is described in the Appendix. It is worth noting that in the case where the voltage is applied, the wave functions are found without using any expansion in terms of the undisturbed wave functions. The numerical computations were performed using the 32-processor computer system Parsytec CC. ## IV The equilibrium state First, we consider the equilibrium state that appears in the absence of an applied voltage $`V_a=0`$. In order to realize the three cases ($`\mathrm{\Delta }\mu >0,\mathrm{\Delta }\mu <0`$, and $`\mathrm{\Delta }\mu =0`$) described in Sec. II, we vary the density of the positive background charge $`n_b`$. In doing this it is convenient to compare $`n_b`$ with the characteristic density $$n_0=\frac{2}{\pi \mathrm{}}\sqrt{2m(\mu _0U_0)},$$ (8) where $`\mu _0`$ is the equilibrium level of the chemical potential in the system. This quantity has a simple physical meaning in the case where the exchange and correlation interaction is absent. It is the background density that determines which of the three cases is realized in the Hartree case. If $`n_b=n_0`$, electrons are not redistributed between the QW and the leads in the equilibration process, if $`n_b>n_0`$, electrons flow from the QW to the reservoirs, and if $`n_b<n_0`$, electrons are transferred from the reservoir to the QW. Of course, turning on the exchange and correlation interaction shifts the value of the background density at which electrons are not redistributed. Nevertheless, as a reference, the value $`n_0`$ remains convenient. The electron density distribution $`n(x)`$ for the three cases is illustrated in Figs. 3, 4, and 5. Shown here are also the graphs of the potential energy $`U(x)`$ that includes the confinement energy and the Hartree energy, $$U(x)=U_0+U_H$$ but not the exchange energy. The exchange energy is not included in $`U(x)`$ because it is a functional of $`\psi `$ rather than a direct function of $`x`$. It is useful to note that $`U_H`$ essentially coincides with the electric potential. Let us consider first the case where the exchange interaction is not taken into account. It is illustrated by the dashed lines in Figs. 3, 4, 5. In this case $`U(x)`$ gives the full single-particle potential shape in the QW. If $`n_b>n_0`$ (this corresponds to $`\mathrm{\Delta }\mu >0`$), Fig. 3 shows that the potential shape lies below the $`U_0`$ energy and hence the interaction energy is negative. This means that a positive charge is accumulated in the QW. It is responsible for the appearance of a potential well. The case $`\mathrm{\Delta }\mu <0`$ is realized when $`n_b<n_0`$. The electron density distribution and the potential shape are shown in Fig. 4. The interaction energy is seen to be positive and the potential shape in the QW lies above $`U_0`$. This means that a negative charge is accumulated in the QW. It produces a potential barrier that hinders electrons in passing through the QW. If $`n_b=n_0`$, the electron density is essentially not redistributed between the QW and the reservoirs. However, Friedel oscillations of the electron density appear near the contacts, Fig. 5. The Friedel oscillations are also observed if there is carrier transfer, but they are superimposed on the much stronger variation of the potential due to the charging of the QW. Let us now consider the exchange interaction effect. The electron density distribution and the potential shape, calculated by taking into account the exchange interaction, are shown in Figs. 3, 4, 5 as solid lines. The exchange interaction is seen to result in an increase of the electron density. This is a consequence of the fact mentioned in Sec. II that the exchange interaction decreases the chemical potential in a QW. That is why more electrons come into the QW when the equilibrium state is established. Correspondingly, the negative charge in the QW increases, which results in the growth of the energy $`U(x)`$. However, this does not yet mean that the electron states with energy lower than $`U(x)`$ are necessarily states decaying in the QW, since the exchange interaction lowers the effective barrier between the QW and the reservoir. Since it would be incorrect to consider the exchange energy as a function of $`x`$, we calculate an average value of the exchange energy per particle incident on the QW with the energy $`\epsilon (k_+)`$ from the left reservoir, $$E_{\mathrm{ex}}=\frac{\psi |\widehat{H}_{\mathrm{ex}}|\psi }{\psi |\psi },$$ (here $`\mathrm{}`$ denotes averaging over the QW length). The average exchange energy $`E_{\mathrm{ex}}`$ is shown in Fig. 6 as a function of $`\epsilon (k_+)`$. The average Hartree energy $`E_H`$ is also given in this figure. It is seen that $`E_{\mathrm{ex}}`$ exceeds $`E_H`$ for all energies. Hence the joint effect of the exchange interaction and the Hartree interaction consists in an effective lowering of the barrier, so that electrons with energy below the confinement energy $`U_0`$ can transit through the QW without decay. Another effect produced by the exchange interaction is also seen from Figs. 3, 4, 5. The exchange interaction strongly enhances the Friedel oscillation amplitude. This result agrees qualitatively with the analytical calculation of the interaction effect on the transmission through a barrier in 1D systems. In our case the Friedel oscillations are generated at the contacts of the QW with the reservoirs. Their amplitude in the potential energy is quite pronounced but smaller than the Fermi energy $`\mu _0U_0`$ in the QW. For the discussion which follows it is important to remark that the Friedel oscillations are superimposed on the smooth variation of the potential produced by the charge accumulated in the QW. Even if the exchange interaction is fully taken into account, this smooth component has an amplitude that is larger than that of the Friedel oscillations. Due to the smooth variation of the potential the QW becomes nonuniform. ## V Linear conductance The model that we have developed above allows us to find the electric current arising when an external voltage is applied. The current is calculated as the sum of the partial currents of the states $`\psi _{r,k_r,s}`$, taking into account their occupation. The linear regime is realized when $`eV_a(\mu _0U_0)`$. In this case we have obtained the dc conductance as a function of the chemical potential $`\mu _0`$. The results of these calculations for zero temperature and for several densities of the background charge $`n_b`$ are given in Fig. 7. The conductance oscillations with varying chemical potential are a consequence of the nonadiabatic reservoir-wire interface. The rapid variation of the potential at this interface leads to backscattering and, if the electron wave is coherent over the entire wire length, to resonances. The oscillations have the same origin as the resonances observed in over the barrier transmission of noninteracting particles. A similar effect also appears in the transport of noninteracting electrons through a narrow, ballistic, nonadiabatic constriction in a 2D electron gas. Our calculations show that the e-e interaction changes the effective potential barrier that electrons have to overcome in passing from one reservoir to another. First, we discuss the results obtained within the Hartree approximation, when exchange interaction is neglected. These results are represented by the lines 1-3 in Fig. 7 for several different background densities $`n_b`$. With increasing $`n_b`$ the oscillations become more frequent, which means that the effective wave number of the electrons is increased. Exactly the same behavior is demonstrated in Figs. 3, 4, 5, namely, with increasing $`n_b`$ the potential in the QW is shifted downward causing the kinetic energy to increase. A similar effect occurs when the exchange interaction is turned on. It is demonstrated in Fig. 7 by curve 3 (obtained by ignoring the exchange interaction) and by line 4 (obtained by including exchange interaction), the background charge being the same in both cases. The exchange interaction is seen to make the conductance oscillations more frequent. The reason for this effect is that the exchange interaction results in an effective lowering of the potential energy of the electrons and consequently in an increase of their kinetic energy. In order to assess the exchange interaction effect on the effective potential, it is instructive to see how the exchange interaction, affects the spectral density of electrons, i. e., $`|\psi _k|^2`$ integrated over the QW length. This is illustrated in Fig. 8. The exchange interaction allows the electrons with energy below the confinement energy $`U_0`$ to pass through the QW. It is interesting to note that, despite the fact that Friedel oscillations are present in the QW, no noticeable suppression of the conductance is observed. The effect of conductance suppression by a periodic potential associated with Friedel oscillations was considered for infinite 1D systems with a $`\delta `$ potential in Refs and . This phenomenon is connected with the fact that a periodic component of the potential suppresses the transmission of the electrons with energy near the Fermi level across the QW (a gap appears at the Fermi level). The absence of this effect in our system is a consequence of two facts. First, the QW has a finite length. Second (and no less essential), the QW becomes inhomogeneous owing to the electron density redistribution between the QW and the leads. As a consequence the kinetic energy at the chemical potential level and the Friedel oscillation period become dependent on the position in the QW. This is why the resonant interaction of electrons at the chemical potential level with the Friedel oscillations is destroyed and the electron passage is not suppressed. ## VI Nonlinear transport and multistability A significant redistribution of the electron density between the QW and the reservoirs occurs under far from equilibrium conditions when the applied voltage exceeds the Fermi energy. Electrons are injected from the left reservoir (cathode) while the electrons entering the QW from the positive reservoir (anode) are scattered back inside the QW. As a consequence, the electron density decreases in the QW (roughly speaking to one-half of the equilibrium density) though the positive background charge is unchanged. Because the positive charge is dominant, a potential well appears in the QW, with the potential shape being distorted by the external potential, as illustrated in Fig. 2. Therefore the kinetic and potential energies are greatly changed. The change in the potential energy produces variations in the wave functions (including even a possibility for resonant states to appear) and the electron density distribution. In this way feedback arises between the electron density and the potential in the QW, which is an important mechanism in nonlinear transport. It is that mechanism which is realized in the model proposed. A complete numerical analysis of the nonlinear transport properties in the wide range of applied voltages within the Hartree-Fock approximation meets some difficulties caused by the long computation time. In this paper we restrict our consideration to the Hartree approximation, which is reasonable at high enough voltage because the exchange energy decreases when the kinetic energy of electrons is increased. The calculations were carried out using the method of pseudotime evolution to the steady solution described in the Appendix. It turns out that in some range of applied voltage an instability of the evolution process appears. The instability origin is not connected with the computation process but is caused by real behavior of the system. The mechanism of the instability is as follows. When the applied voltage is high enough (compared to the Fermi energy), the electron flow injected from the negatively charged reservoir is the only flow in the QW. Let a velocity fluctuation appear in some portion of the wire. To be definite, let us assume that the velocity is increased above its stationary value. Since the total electron flow is limited by the contact, it is not disturbed by this fluctuation. Hence, the continuity of the current requires that the electron density decreases. This leads to a growth of the positive (net) charge, because electrons cannot completely neutralize the background charge. The excess positive charge causes the potential energy of the electrons to decrease. Under the condition of ballistic transport, this results in a new increase of the velocity, and so on until some nonlinear process stabilizes this instability. In our model this is achieved by a redistribution of the overall electron density and a reshaping of the potential distribution in the QW. In such a way the potential shape is switched from one state to the other under the condition that both states are characterized by the same potential difference across the QW ends. In order to describe the transition from one shape to the other as a continuous process it is necessarily to characterize the nonequilibrium state of the system by a parameter other than the applied voltage. This parameter should distinguish states with different potential shapes and the same potential difference across the ends. As such a parameter, the mean kinetic energy $`E_{\mathrm{kin}}`$ of electrons in the QW can be used, $$E_{\mathrm{kin}}=\frac{_{r=\pm }𝑑k_rk_r|\widehat{T}|k_r}{_{L/2}^{L/2}𝑑xn(x)},$$ where $`\widehat{T}`$ is the kinetic energy operator. This conclusion is similar to what is known in the theory of hot electron instabilities in semiconductors. The electron heating by the electric field results in S- and N-shaped current-voltage characteristics. Under these conditions, the electron temperature uniquely determines the state of the system. In our case the kinetic energy of the electrons is a direct analog of the electron temperature. We have developed the algorithm that allows one to solve our problem in the case where the mean kinetic energy is fixed rather than the applied voltage. This algorithm is described in the Appendix. In this computation scheme all quantities (including $`V_a`$) are determined by $`E_{\mathrm{kin}}`$. We have found that this algorithm gives stable results and the nonequilibrium state of the system under investigation is uniquely determined by $`E_{\mathrm{kin}}`$. In particular, the dependence of $`E_{\mathrm{kin}}`$ on the applied voltage is shown in Fig. 9. The kinetic energy is seen to have several values for a given voltage $`V_a`$, while $`V_a`$ is uniquely defined by $`E_{\mathrm{kin}}`$. Correspondingly, several nonequilibrium states, with different distributions of the electron density and the potential, are possible at a given voltage. It is obvious that the states are not all stable with respect to time-dependent fluctuations. The multistability, and in particular bistability, phenomenon described above may be useful for understanding the negative differential conductance observed in quantum wires. ## VII Conclusions In this paper a QW of a finite length coupled to reservoirs is considered as a unified system. We have found that the electron density is substantially redistributed between the wire and the electron reservoirs when this system is formed. As a consequence of this process, a QW can acquire a net charge. The charging of the wire is caused by the chemical potential difference between the wire and the leads which exists if these two subsystems are decoupled. This phenomenon is similar to the contact potential difference in a classical multiconductor system. The structures of the charge density and the electric potential distributions differ according to the chemical potential difference: (i) a positive charge is accumulated in the wire and a potential well is developed there; (ii) the wire is charged negatively giving rise to a potential barrier; (iii) the wire remains uncharged as a whole. In all cases, Friedel oscillations are present which are generated at the nonadiabatic contacts of the QW with reservoirs. They are superimposed on the relatively smooth profile of the potential produced by the charge accumulated in the wire. This smooth potential has a large amplitude. The Friedel oscillation amplitude is strongly enhanced if the exchange interaction is included. Variation of the electron density in a QW due to electron redistribution between the wire and the reservoirs produces a significant effect on dc conductance. This effect is connected with the change of the kinetic energy of the electrons due to two factors: (i) the variation of the accumulated charge and the potential variation associated with this charge; (ii) the variation of the exchange energy. The exchange energy substantially lowers the effective potential barrier that electrons have to overcome when passing from one reservoir to another. The Friedel oscillation potential does not suppress the conductance because the QW becomes inhomogeneous as a consequence of the electron density redistribution between the QW and the leads. The electron density variation due to the voltage applied across the leads and the change of the effective potential shape, associated with this variation, in the wire is an important mechanism for nonlinear transport. The most interesting consequence of the charging effect in the wire is the instability that arises under a high enough applied voltage. The instability shows itself as a spontaneous increase of the kinetic energy of the injected electrons at a given applied voltage. In a QW with a nonadiabatic connection to the reservoirs, the development of instabilities results in the appearance of multistable states, i.e., in the existence of several stable states at a given voltage. We conclude by emphasizing that the charging effect analyzed here is a general phenomenon that might be important in many mesoscopic systems containing electronically different compounds. Examples of current interest are carbon nanotubes, hybrid normal-superconducting systems, and atomic quantum point contacts. ## Acknowledgments The present work has been supported by INTAS (Grant No. 96-0721). V.A.S. and S.V.P acknowledge the Russian Fund for Basic Research (Grant No. 99-02-18192), the Russian Program ”Physics of Solid-State Nanostructures” (Grant No. 97-1054) and the Russian Program ”Surface Atomic Structures” (Grant No. 5.3.99). M.B. is funded by the Swiss National Science Foundation. The multiprocessor computer system Parsytec CC used for the numerical calculations has been acquired by the Institute for Mathematical Modeling of the Russian Academy of Sciences with financial support from the EC ESPRIT program (Grant No. 21041). ## A The computation method The problem of finding the wave functions for a QW coupled to reservoirs can be reduced to the solution of Eq. (5) in the inner region $`L/2xL/2`$ and matching $`\psi _{r,k,s}`$ and its derivative at the boundaries $`x=\pm L/2`$ to the wave functions (4) in the outer regions. Combining the matching equations for $`\psi _{r,k,s}`$ and $`\psi _{r,k,s}^{}`$, one can exclude the coefficients $`R_r`$ and $`T_r`$ to get finally the following nonuniform boundary conditions for $`\psi _{r,k,s}`$: $$\psi _{r,k_r,s}^{}=\{\begin{array}{ccc}ik_r(2\psi _{r,k_r,s}),\hfill & & rx=L/2\hfill \\ ik_r^{}\psi _{r,k_r,s},\hfill & & rx=L/2,\hfill \end{array},$$ (A.1) where $`k_r`$ and $`k_r^{}`$ are defined in Eq. (4) The wave functions $`\psi _{r,k_r,s}`$ are considered as functions of two continuous variables: the space coordinate $`x`$ and the energy $`\epsilon _r=\mathrm{}^2k_r^2/2m`$. Thus $`\psi _{r,k_r,s}=\psi _{r,s}(x,\epsilon _r)`$, where $`x`$ and $`\epsilon _r`$ are varied respectively in the regions $`L/2xL/2`$ and $`0<\epsilon _r\epsilon _M`$, with the upper boundary $`\epsilon _M`$ being equal to $`\mu _0+3k_BT`$ ($`k_B`$ is the Bolzmann constant, $`T`$ is the temperature). In this region the uniform grid $`\{x_i,\epsilon _j\}_{i=0,N_1;j=0,N_2}`$ is constructed. The equation for $`\psi _{r,k,s}(x_i,\epsilon _j)`$ on the grid is obtained from Eq. (5) with use of the integro-interpolative method for the node presentation on $`x_i`$ and the trapezium formula when calculating the integrals on $`k_r`$ for the nodes on $`\epsilon _j`$. This results in a set of nonlinear finite-difference equations that can be symbolically presented in the form $$\widehat{𝐌}[𝚿]𝚿=𝐅,$$ (A.2) where $`𝚿`$ is the wave function vector to be found and $`\widehat{𝐌}[𝚿]`$ is the nonlinear operator. The matrix equation (A.2) is inhomogeneous as a consequence of the boundary conditions (A.1). The equations (A.2) are solved by the iteration method. However, the commonly used successive approximation scheme ($`\widehat{𝐌}[𝚿^{(l)}]𝚿^{(l+1)}=𝐅`$, with $`l`$ being the iteration number) turns out to be badly convergent. We use the method of pseudotime evolution to the steady solution. More specifically, we use the two-layer iteration scheme of this method. In this scheme the approximating matrix $`\widehat{𝐌}^{(l)}`$ is introduced, which is calculated with the use of the iteration process $$\frac{\widehat{𝐌}^{(l+1)}\widehat{𝐌}^{(l)}}{\tau _l}=\widehat{𝐌}[\mathrm{\Psi }^{(l)}]\widehat{𝐌}^{(l)},l=0,1,2,\mathrm{},$$ (A.3) where $`\tau _l`$ is a pseudotime parameter. The choice of $`\tau _l`$ allows one to attain the best convergence of the iteration process. As the starting value of $`\widehat{𝐌}^{(0)}`$ we use $`\widehat{𝐌}[\mathrm{\Psi }=0]`$, i.e., the $`\widehat{𝐌}`$ matrix for noninteracting electrons. During the iteration process, $`\mathrm{\Psi }^{(l)}`$ is calculated with the use of the equation $$\widehat{𝐌}^{(l)}𝚿^{(l)}=𝐅.$$ The pseudotime $`\tau _l`$ is determined by the $`\widehat{𝐌}^{(l)}`$ operator spectrum. The optimal convergence is attained when $$\tau _l=\frac{2}{\lambda _{\mathrm{min}}(M^{(l)})+\lambda _{\mathrm{max}}(M^{(l)})},$$ where $`\lambda _{\mathrm{min}}`$ and $`\lambda _{\mathrm{max}}`$ are the lowest and highest eigenvalues of $`\widehat{𝐌}^{(l)}`$. The iteration process is ended when the following condition is fulfilled $$\mathrm{max}\left|\frac{m_{ij}^{(l+1)}m_{ij}^{(l)}}{m_{ij}^{(l)}}\right|\delta ,$$ where $`m_{ij}^{(l)}`$ is an element of the $`𝐌^{(l)}`$ matrix. In the present paper, $`\delta `$ was chosen to be $`10^6`$. The above method is successful when the system under investigation has a unique solution. However, at some fixed values of the applied voltage the computation shows an instability. In the course of the pseudotime evolution process the calculated quantities (such as the potential, the electron density, the kinetic energy) are randomly switched between several values. This is connected with the fact that the state of the system is not uniquely determined by the calculation scheme where the applied voltage is fixed. A unique description of the system is achieved by using the mean kinetic energy $`E_{\mathrm{kin}}`$ of the electrons as the parameter that defines the nonequilibrium state of the system. We have developed computation algorithm that allows one to vary $`E_{\mathrm{kin}}`$ continuously, in other words, we solve the problem using $`E_{\mathrm{kin}}`$ as the fixed parameter instead of the applied voltage. An essential question appearing in this algorithm is how the applied voltage $`V_a`$ should be defined when $`E_{\mathrm{kin}}`$ is given. The equation defining $`V_a`$ is obtained from Eq. (5). Multiplying this equation by $`\psi _{r,k_r,s}^{}`$, integrating it over $`k_r`$ and over $`x`$, and summing over $`r`$ one gets an equation of the following form: $$E_{\mathrm{kin}}eV_aA[𝚿]=B[𝚿],$$ (A.4) where $`A[𝚿]`$ and $`B[𝚿]`$ are functionals of the electron wave functions. Solving this equation with respect to $`V_a`$ one gets $`V_a`$ as a functional of $`𝚿`$, with $`E_{\mathrm{kin}}`$ being a parameter, $$V_a=\mathrm{\Phi }_{E_{\mathrm{kin}}}[𝚿].$$ (A.5) When solving the problem with $`E_{\mathrm{kin}}`$ as a parameter, Eq. (A.5) should be taken into account together with Eq. (A.3). This system of equations is solved using the above pseudotime evolution method and two-layer iteration scheme. The set $`V_a^{(l)}`$ approximating $`V_a`$ is defined as $$\frac{V_a^{(l+1)}V_a^{(l)}}{\tau _l}=\mathrm{\Phi }_{E_{\mathrm{kin}}}[𝚿^{(l)}]V_a^{(l)},l=0,1,2,\mathrm{}.$$ As the starting value of the $`V_a^{(l)}`$ set, the arbitrary value of $`V_a`$ in the stability region close to the instability threshold can be used. In this generalized procedure the pseudotime $`\tau _l`$ is chosen taking into account the spectral properties of the total matrix $`𝐌^{(l)}V_a^{(l)}`$.
no-problem/9908/cond-mat9908015.html
ar5iv
text
# Angle-resolved photoemission study of untwinned PrBa2Cu3O7: undoped CuO2 plane and doped CuO3 chain ## I Introduction In the study of YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> (YBCO) and its family cuprates, the physical properties of the CuO<sub>3</sub> chain itself have been a subject of interest as well as those of the CuO<sub>2</sub> plane. In particular, evidence for charge instability has been observed in the hole-doped CuO<sub>3</sub> chain of YBCO and its relation with superconductivity has been discussed. Among the YBCO family compounds, PrBa<sub>2</sub>Cu<sub>3</sub>O<sub>7</sub> (PBCO) is unique in that it does not show superconductivity while the other rare-earth substituted YBCO compounds are superconducting. Optical studies have revealed that the CuO<sub>2</sub> plane is not doped with holes and consequently that the superconductivity is suppressed in PBCO. In order to explain the hole depletion in the CuO<sub>2</sub> plane, several models have been proposed. Among them, the model proposed by Fehrenbacher and Rice (FR), in which the Pr 4$`f`$-O 2$`p_\pi `$ states trap holes from the Cu 3$`d`$-O 2$`p_\sigma `$ band, has been most successful. It was also argued that hole transfer between the CuO<sub>2</sub> plane and the CuO<sub>3</sub> chain may play an important role. Experimental studies of the band structure of the CuO<sub>2</sub> plane, the CuO<sub>3</sub> chain and the Pr 4$`f`$ state should further reveal the difference between the superconducting YBCO and non-superconducting PBCO. Another important and potentially even more interesting point is that, if the CuO<sub>2</sub> plane is undoped and the CuO<sub>3</sub> chain is doped as suggested by the various experimental and theoretical studies, PBCO may give us an opportunity to study a hole doped CuO<sub>3</sub> chain with good one-dimensionality compared to YBCO. Recently, an angle-resolved photoemission spectroscopy (ARPES) study of SrCuO<sub>2</sub> by Kim et al. has shown the spinon and holon bands of the undoped CuO<sub>3</sub> chain, which is a manifestation of spin-charge separation in the one-dimensional system, and has attracted much interest. However, no ARPES study of a hole-doped CuO<sub>3</sub> chain has been made so far. Although YBCO has been studied by ARPES, the band dispersion from the CuO<sub>3</sub> chain is far from being one dimensional because of the strong interaction between the chain and the plane. In this paper, we report an ARPES study of untwinned PBCO samples which have relatively low resistivity but is not superconducting. As we show below, we observe at least two kinds of dispersive features: one from the undoped antiferromagnetic CuO<sub>2</sub> plane and the other which is highly one-dimensional band from the doped CuO<sub>3</sub> chain. ## II Experimental Single crystals of PBCO were grown in a MgO crucible by a pulling technique. The composition was determined to be Pr<sub>1.02</sub>Ba<sub>1.98</sub>Cu<sub>2.92</sub>Mg<sub>0.08</sub>O<sub>7</sub> by inductively coupled plasma analysis. Because a small amount of Mg-impurities originating from the crucible are substituted preferentially for the plane Cu sites, it is expected that the CuO<sub>3</sub> chain is not affected by these impurities. Rectangular shaped samples were cut out from the as-grown crystal and annealed at 500 C in oxygen atmosphere under uniaxial pressure. The resistivity parallel to the chain direction increases with cooling from 8 m$`\mathrm{\Omega }`$cm at 300 K up to 10 $`\mathrm{\Omega }`$cm at 10K and that perpendicular to the chain direction increases from 400 m$`\mathrm{\Omega }`$cm at 300 K to 70 $`\mathrm{\Omega }`$cm at 10K. The ARPES measurements were performed using the Vacuum Science Workshop chamber attached to the undulator beamline 5-3 of Stanford Synchrotron Radiation Laboratory (SSRL). Incident photons were linearly polarized and had energy of 29 eV. The total energy resolution including the monochromator and the analyzer was approximately 40 meV. The angular resolution was $`\pm 1`$ degree, which gives the momentum resolution of $`\pm `$0.05$`\pi `$ at $`h\nu `$ = 29 eV. \[In the text, momenta along the $`a`$\- and $`b`$-axes are given in units of $`1/a`$ and $`1/b`$, respectively. Here, $`a`$ (= 3.87 $`\stackrel{}{\mathrm{A}}`$) and $`b`$ (= 3.93 $`\stackrel{}{\mathrm{A}}`$) are the lattice constants of PBCO perpendicular and parallel to the chain direction, respectively.\] The chamber pressure during the measurements was less than $`5\times 10^{11}`$ Torr. The samples were cooled to 10 K and cleaved in situ. The cleaved surface was the $`ab`$ plane. Orientation of the $`a`$\- and $`b`$-axes was done by Laue diffraction before and after the measurement. The cleanliness of the surfaces was checked by the absence of a hump at energy $`9.5`$ eV. All the spectra presented here were taken within 30 hours of cleaving. We cleaved the samples three times and checked the reproducibility. As shown in Fig. 1, ARPES data were taken in two arrangements: an $`Eac`$ arrangement in which the photon polarization $`E`$ is in the $`ac`$ plane and is perpendicular to the chain and an $`Ebc`$ arrangement in which $`E`$ is in the $`bc`$ plane and has a component parallel to the chain. The position of the Fermi level ($`E_\mathrm{F}`$) was calibrated with gold spectra for every measurement. ## III Results and discussion The entire valence band data measured at ($`k_a`$,$`k_b`$) = (0,0) are shown in Fig. 2. Here, $`k_a`$ and $`k_b`$ are momenta along the $`a`$\- and $`b`$-axes, respectively. The Ba 5$`p`$ core levels are split into surface and bulk components as observed in YBCO, indicating that the cleaved surface is as good as that of YBCO. The peak labeled as A is intense for $`Ebc`$, while it loses its weight for $`Eac`$. This strong polarization dependence indicates that peak A is derived from the CuO<sub>3</sub> chain and that the CuO<sub>3</sub> chain is well aligned at the surface. Peak A can be attributed to the non-bonding O 2$`p`$ states of the CuO<sub>3</sub> chain. On the other hand, the intensity of peak B is insensitive to the photon polarization, indicating that it is derived from the CuO<sub>2</sub> plane. Peak B has energy of $``$ -2.3 eV and is similar to that found in Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub>. As pointed out by Pothuizen et al., peak B corresponds to the non-bonding O 2$`p`$ states of the CuO<sub>2</sub> plane. Figure 3 shows ARPES spectra along the (0,0)$``$($`\pi `$,$`\pi `$) direction and those along the (0,$`\pi `$/2)$``$($`\pi `$,$`\pi `$/2) direction taken with $`Eac`$. The uppermost and second uppermost spectra in the left panel were taken at (0,0) and ($`\pi `$/4,$`\pi `$/4), respectively. The intensity from -0.2 to -0.4 eV is enhanced at ($`\pi `$/4,$`\pi `$/4) compared to that at (0,0). Let us denote this feature as $`\alpha `$ and discuss it in following paragraphs. The spectra taken at ($`\pi `$/2,$`\pi `$/2) are shown by the thicker solid curves and marked by the closed circles in the two panels of Fig. 3. Figure 4 shows ARPES spectra nearly along the (0,0)$``$($`\pi `$,$`\pi `$) direction for $`Ebc`$. In Fig. 4, the spectrum taken at ($`\pi `$/2,$`\pi `$/2) is shown by the thicker solid curve and marked by the closed circle. In these spectra taken around ($`\pi `$/2,$`\pi `$/2), a dispersive feature with a band maximum at $``$ -0.4 eV is clearly seen and is labeled as $`\beta `$. This structure $`\beta `$ is very similar to that found at ($`\pi `$/2,$`\pi `$/2) in Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub> and can be interpreted as the Zhang-Rice (ZR) singlet state of the undoped antiferromagnetic CuO<sub>2</sub> plane. The energy difference between the ZR state at ($`\pi `$/2,$`\pi `$/2) (structure $`\beta `$) and the non-bonding O 2$`p`$ state at (0,0) (peak B) is $``$ 2 eV in PBCO, in agreement with the observation in Sr<sub>2</sub>CuO<sub>2</sub>Cl<sub>2</sub>. In addition to the ZR state, a weak structure $`\gamma `$ at -0.2 eV was observed around ($`\pi `$/2,$`\pi `$/2). By comparing Figs. 3 and 4, one can notice that the intensity of structure $`\alpha `$ at ($`\pi `$/4,$`\pi `$/4) is weak for $`Ebc`$ compared to that for $`Eac`$. In addition, while the relative intensity of structure $`\alpha `$ to structure $`\beta `$ has strong polarization dependence, that of structure $`\gamma `$ to structure $`\beta `$ has only small polarization dependence. This suggests that these two structures $`\alpha `$ and $`\gamma `$ have different origins. As discussed in the next paragraph, structure $`\alpha `$ at ($`\pi `$/4,$`\pi `$/4), which shows strong polarization dependence, is part of a one-dimensional band from the CuO<sub>3</sub> chain. On the other hand, since structure $`\gamma `$ shows weak polarization dependence and probably has two-dimensional character, it is tempting to interpret structure $`\gamma `$ as a Pr 4$`f`$-O 2$`p_\pi `$ hybridized state or a so-called FR state. If this is the FR state, the present spectra are consistent with the FR scenario because the FR state is closer to $`E_\mathrm{F}`$ than the ZR state of the CuO<sub>2</sub> plane. The situation that the FR state is partially occupied and is observable in photoemission spectroscopy is consistent with the optical result, which has shown that the formal valence of the Pr ion is +3.5 and the FR state is occupied by 0.5 electrons on the average if it exists. This argument on structure $`\gamma `$ should be confirmed by using Pr 4$`d`$-4$`f`$ resonant photoemission in the future. In Figs. 5 and 6, we have plotted ARPES data along the chain direction taken with $`Eac`$. In the spectra taken at ($`k_a`$,0) ($`k_a`$ is 0, $`\pi `$/10, $`\pi `$/5, 3$`\pi `$/10, or $`\pi `$), which are shown at the uppermost position in each panel, the broad feature labeled as $`\alpha ^{}`$ is located at -0.7 $`\pm `$ 0.2 eV. This feature moves toward $`E_\mathrm{F}`$ as $`k_b`$ increases and reaches a band maximum with energy of $``$ -0.2 eV around ($`k_a`$,$`\pi `$/4). The spectra taken at ($`k_a`$,$`\pi `$/4) are shown by the thicker solid curves and marked by the closed circles in each panel of Figs. 5 and 6. This feature loses its weight for $`k_b>\pi /4`$ and the spectrum at ($`k_a`$,$`\pi `$/2) is almost featureless. This behavior does not depend on $`k_a`$, i.e., the momentum perpendicular to the chain, meaning that this dispersive feature is highly one dimensional. In order to demonstrate the dispersion more clearly, the contour plots of the difference spectra, which are obtained by subtracting the featureless spectrum at ($`\pi `$,$`\pi `$/2) from the spectra at each $`k`$-point, are shown for $`k_a`$ = 0, 3$`\pi `$/10, and $`\pi `$ in Fig. 7. Indeed, the overall dispersion of structure $`\alpha ^{}`$ does not depend on $`k_a`$ so much. Therefore, we can conclude that the band is derived from the CuO<sub>3</sub> chain. Structure $`\alpha `$ at ($`\pi `$/4,$`\pi `$/4) \[see Fig. 3\] is also part of this one-dimensional band. The fact that the one-dimensional band reaches a band maximum at $`k_b`$ $``$ $`\pi `$/4 and disappears for $`k_b>\pi /4`$ indicates that the filling of the Cu 3$`d_{x^2y^2}`$-O 2$`p_\sigma `$ band in the CuO<sub>3</sub> chain is close to 1/4, namely, the formal valence of Cu in the chain is $``$ +2.5. This is consistent with the optical study. There are two interesting points in this one-dimensional band. The first point is that the other structure ($`\alpha ^{\prime \prime }`$), which has higher energy than structure $`\alpha ^{}`$, is observed around ($`\pi `$,0) \[Compare the uppermost spectra shown by the thicker solid curves in the two panels of Fig. 5.\]. In order to show the dispersion clearly, the density plots of the second derivatives of the ARPES spectra along the chain direction are diplayed for $`k_a`$ = 0, 3$`\pi `$/10, and $`\pi `$ in Fig. 8. The dispersion of $`\alpha ^{\prime \prime }`$ is visible in the spectra for $`k_a`$ = $`\pi `$ which is shown in the right panel of Fig. 8. It is possible to attribute $`\alpha ^{}`$ and $`\alpha ^{\prime \prime }`$ to the holon and spinon bands of the Tomonaga-Luttinger (TL) liquid. While the total width of the holon band is predicted to be $``$ 4$`t`$, that of the spinon band has an energy scale of $`J`$. In Figs. 7 and 8, model holon and spinon dispersions of -2$`t`$cos($`k_b`$+$`\pi `$/4) and -$`\pi J/2`$cos(2$`k_b`$) are shown by solid curves. The two curves with $`t`$ of 0.5 eV and $`J`$ of 0.16 eV, which are reasonable values for the cuprates, roughly follow the dispersions in Fig. 8. The second point is that the spectral weight near $`E_\mathrm{F}`$ is considerably suppressed in the experimental data in disagreement with the theoretical prediction on the TL liquid. Although the lengths of the CuO<sub>3</sub> chain at the surface are finite because of the surface termination, the observation of the nice dispersive behavior indicates that the lengths of the CuO<sub>3</sub> chains are long enough to allow us to compare the data with the theory for the doped CuO<sub>3</sub> chain. A possible origin of the intensity suppression near $`E_\mathrm{F}`$ is the instability of the nearly-1/4-filled CuO<sub>3</sub> chain leading to charge density waves (CDW). Actually, charge instability in the hole-doped CuO<sub>3</sub> chain of PBCO has been observed by NMR and NQR measurements. In addition, it has recently been pointed out that the spectral function of one-dimensional CDW insulators can have the holon and spinon dispersions. Here, it should be remarked how we would observe the feature with $`Eac`$. In PBCO, the CuO<sub>4</sub> square plane of the CuO<sub>3</sub> chain is perpendicular to the $`ab`$ plane, i.e., the sample surface. Therefore, the photon polarization has a component perpendicular to the sample surface, namely, parallel to the CuO<sub>4</sub> square plane of the CuO<sub>3</sub> chain as shown in Fig. 1. It is this additional component of the polarization that gives a finite transition matrix element to the ZR state in the CuO<sub>3</sub> chain. With $`Ebc`$, the contribution from the chain is very weak when $`k_a`$ is small. As $`k_a`$ becomes larger, the intensity of the one-dimensional band for $`Ebc`$ becomes larger and, at $`k_a`$ = $`\pi `$, is comparable to that for $`Eac`$. Figure 9 shows ARPES spectra along the ($`\pi `$,$`\pi `$/2)$``$($`\pi `$,-$`\pi `$/2) direction, namely, along the chain direction taken with $`Ebc`$. The two dispersive features, which can be interpreted as the holon and spinon dispersions, are also observed for $`Ebc`$. In order to show the dispersion obtained for $`Ebc`$ clearly, the density plot of the second derivatives of the ARPES spectra is displayed in Fig. 10. The dispersions of the holon and spinon bands are almost symmetric with respect to $`k_b`$ = 0 and reach maxima around $`k_b`$ = $`\pi `$/4 and -$`\pi `$/4. These dispersions obtained for $`Ebc`$ are consistent with those obtained for $`Eac`$. In the present ARPES data taken with $`Eac`$ and $`Ebc`$, the relative intensity of the spinon band to the holon band strongly depends on $`k_a`$. The spinon band becomes more intense as $`k_a`$ becomes larger. The same behavior was found in the undoped CuO<sub>3</sub> chain by Kim et al. Further experimental and theoretical investigation is required to reveal this peculiar $`k_a`$ dependence of the spectral function. ## IV Conclusion In conclusion, we have observed the band dispersions from the CuO<sub>2</sub> plane and the hole-doped CuO<sub>3</sub> chain of non-superconducting PBCO. The band dispersion from the CuO<sub>2</sub> plane clearly shows that PBCO has an undoped insulating CuO<sub>2</sub> plane. On the other hand, the one-dimensional dispersive feature from the doped CuO<sub>3</sub> chain consists of two structures which can be interpreted as holon and spinon bands. These bands lose intensity beyond $`k_b`$ $``$ $`\pi `$/4, indicating that the CuO<sub>3</sub> chain is nearly 1/4-filled. Further experimental and theoretical studies are desirable to reveal the nature of the one-dimensional band including its momentum and polarization dependence. ## Acknowledgment We would like to thank M. Schabel, I. Terasaki, T. Thoyama, S. Maekawa, D. D. Sarma, and K. Penc for valuable comments. We are grateful to P. J. White, A. Y. Matsuura and the staff of SSRL for technical support. This work was supported by a Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture of Japan, Special Coordination Funds of the Science and Technology Agency of Japan, the New Energy and Industrial Technology Development Organization (NEDO), the U. S. DOE, Office of Basic Energy Science and Division of Material Science. SSRL is operated by the U. S. DOE, Office of Basic Energy Sciences, Division of Chemical Sciences.
no-problem/9908/gr-qc9908031.html
ar5iv
text
# On the Nature of Black Hole Entropy ## Black hole entropy and internal states Let me begin by giving several reasons why we should not think that the Bekenstein-Hawking entropy $`S_{\mathrm{BH}}=A/4\mathrm{}G`$ of a black hole counts the number of internal states of the black hole. (By “the entropy” of a black hole I will always mean $`A/4\mathrm{}G`$ in this article.) These reasons have been enunciated in a thoughtful article by Rafael Sorkinsorksm , which I will borrow from here. 1. The spatial region inside a black hole horizon can have arbitrarily large volume, with room for an arbitrarily large number of states. For example a Friedmann universe of any size can be joined to the interior of a Schwarzschild black hole. Thus the number of possible internal states of a black hole is unbounded. 2. A black hole is not in “internal equilibrium”, so why should its thermodynamic entropy refer to its interior states? 3. Conditions inside the horizon are causally disconnected from the outside, so how can the states inside be thermodynamically relevant to the outside? 4. According to local quantum field theory the evaporation of a black hole is unitary, at least until the final stages, and the Hawking radiation is correlated to field degrees of freedom inside the black hole. The number of internal states of the black hole must therefore remain large enough to store all the correlations maintaining the purity of the total state. As a black hole evaporates, however, its area and therefore its entropy decreases. Thus the entropy must not be counting the number of internal states. Regarding point 1, it should be mentioned that the example given will have a white hole horizon and singularity in its past (assuming the weak energy condition holds) so it is not a configuration that would evolve from an ordinary collapse processsorwaljiu . It is nevertheless a possible state of the black hole. There is by now a “standard” argument against points 3 and 4, namely, that local quantum field theory may be inapplicable. This argument is suggested by (but not restricted to) string theory, in which local quantum field theory is only an approximation valid under certain conditions. It has been argued both on general principlesthooftsmatrix and in string theorynonlocal that there are no truly local observables in quantum gravity and that for this reason the decomposition of the Hilbert space into sectors inside and outside the black hole is invalid from the beginning. While this may indeed be true at some fundamental level, the relevant question here is whether local quantum field theory holds to a sufficiently good approximation for points 3 and 4 to be valid. Since the black hole can be macroscopic and the curvature can be very small compared with the string length or Planck length, it is hard for me to see why the local field theory approximation should fail in this regard. To postulate such a mysterious failure, when simpler scenarios exist, seems to me uncalled for radicalism, although it is a hypothesis favored by many physicists today. ## Black hole entropy and surface states The previous arguments point to the conclusion that black hole entropy is a measure of only those states that can influence the outside of the black hole.<sup>1</sup><sup>1</sup>1The case for a surface interpretation of black hole entropy has been made by various authors. In particular, an article by BanksBanks:1994ph (written before the age of D-branes) makes the case with many of the same arguments as used here, and the argument that the universality of black hole entropy (in spite of the non-universal history of the black hole) arises from the universality of the near-horizon geometry was made in a paper by Parentani and PiranP2 . These states must be associated with the presence of the horizon, otherwise they would simply be counted as ordinary states of the exterior itself. One interpretation of this “surface entropy” is that it measures the information in the entanglement of the vacuum across the horizon (“entanglement entropy”)sorkvac . For fields on a fixed background this is equivalentkabastra to the entropy of the thermal state (“thermal atmosphere”) that results when the state is restricted to the outsidethooftwall . This entropy diverges, but gives something of the correct order of magnitude if a Planck scale cutoff is imposed. It is insufficient to consider fields on a fixed background however. For one thing, although the contributions of quantum fields can be thought of as “loop corrections” to the black hole entropy, there is also a classical contribution coming from the gravitational action itself. On can imagine an induced gravity scenariosakharov ; tajind ; froind1 ; froind2 , in which the entire gravitational action is induced by matter, however there is still another problem: for non-minimally coupled scalar fields or gauge fields, the entanglement entropy is not equal to the corresponding contribution to the entropy computed from the induced gravitational action. It seems that the difference between these two entropies can be understood as a consequence of the fact that the background itself varies when the temperature is variedfrofurzel ; frofur . Physically, this means that to understand the entropy one must count states in the coupled matter-gravity vacuum. The large and universal number of states per unit of surface area seems to be explained by the infinite redshift at the horizon: many states at short distances near the horizon have the same, low, energy. In fact, the number would appear to be infinite from perturbative counting, but the final count requires knowledge of only the low energy effective gravitational action and the associated low energy Newton constant, as long as the spacetime curvature is small compared with Planck curvaturesusugl ; tajind ; larwil . Although we are unable to compute the renormalized Newton constant from quantum gravity, its (finite) value can be measured and used in the entropy formula. ### Entanglement entropy and the generalized second law Sorkin proposed a derivation of the generalized second law based on the entanglement interpretation of black hole entropysorksm ; sorksl . His idea was that the total entropy $`S_{\mathrm{outside}}=S_{\mathrm{horizon}}+S_{\mathrm{rest}}`$ of the reduced density matrix outside the horizon receives a large universal contribution $`S_{\mathrm{horizon}}`$ from the vicinity of the horizon and the rest $`S_{\mathrm{rest}}`$ is primarily just the ordinary entropy of a mixed state outside. Invoking the dynamical autonomy of the evolution outside the horizon, Sorkin argues that $`S_{\mathrm{outside}}`$ cannot decrease, which amounts to the usual generalized second law provided $`S_{\mathrm{horizon}}`$ can be identified with the black hole entropy. This explanation of the generalized second law seems so natural that it is hard to believe there is not some truth in it. Unfortunately, as mentioned above, the entanglement interpretation of black hole entropy does not seem to work, but perhaps this conclusion is premature. Perhaps the black hole entropy could yet be understood in terms of entanglement entropy if, as proposed in barfrozel , the division of the system into inside and outside is referred to an intrinsic feature of the fluctuating geometry such as the minimal throat area on some preferred spacelike slice. ## Objections Objections can be raised to the assertion that black holes have many more states than are counted by the black hole entropy. I believe that all of these objections are wrong, but it is challenging and instructive to try to point to exactly where they are wrong. I will try to do so here with regard to several objections, all but the first coming from string theory. ### Black hole pair creation amplitudes Semiclassical calculations of black hole pair creation rates display a factor $`\mathrm{exp}S_{\mathrm{BH}}`$ which admits the natural interpretation as a density of states factorpaircreation . This seems to lend solid support to the interpretation of $`\mathrm{exp}S_{\mathrm{BH}}`$ as the number of states of the black hole. If the black hole had more states, would they not contribute to the pair creation rate? This question has been discussed in the past, with conflicting conclusionsBanks:1993mi ; Banks:1993is ; Banks:1994ph ; Gid:1994vj ; Gid:1995qt , and it deserves to be discussed further. Here I will only state the reason for my belief that the answer is no<sup>2</sup><sup>2</sup>2I was asked this question during my talk and had no quick answer. After the talk Renaud Parentani suggested the following answer.: Pair creation is an exponentially suppressed tunneling process, and any “unnecessary” decoration of the black holes would, it seems, be even more suppressed. All the extra internal states are unnecessary decoration, and are therefore essentially irrelevant to the pair creation rates. ### String theory Calculations of black hole entropy in string theory and its descendents have been carried out in several contexts yielding agreement with the Bekenstein-Hawking entropy. In all cases it appears that one is indeed counting all of the states of the object identified with a black hole. Can this be compatible with the claim that black hole entropy does not count all of the states? I will attempt to argue that it can, pointing to where one might find the other states. My attempts are only partly successful, and are particularly weak in the context of the AdS/CFT duality. #### D-branes The entropy of certain near-extremal configurations of D-branes has been found to agree with the semiclassical entropy of the black hole configurations with the same set of charges (see for exampleDbranesMalda ; DbranesPeet ). In the extremal case, for the supersymmetric BPS states, this is understood as a consequence of the fact that the D-brane configuration evolves into the black hole as the string coupling is increased from weak to strong, all the while maintaining the supersymmetry. The enumeration of BPS states is independent of the coupling, hence the agreement in the count of states. In the D-brane picture there is nothing corresponding to the inside of the black hole where extra states can reside so, given the agreement with the black hole entropy, how could a black hole have any more states? For the BPS states the answer is simple: the black hole also does not have any interior in the sense that on a spacelike slice orthogonal to the timelike Killing field the horizon is infinitely far away and has no other side. The D-brane and black hole entropies also agree for near-extremal states however. In these cases, one can not give such a simple answer. Imagine for instance a configuration that has been maintained at fixed energy above extremality for a long time with the help of an influx of energy equal in magnitude to the Hawking flux. In the black hole picture there is an arbitrarily large amount of information stored in the correlations between the inside and outside of the black hole, so there must be a correspondingly large number of states for the interior. In the D-brane description however there is nothing that corresponds to the interior. How could there be such a drastic mismatch between the total number of states in the two descriptions and still be such agreement on not only the entropy but also the rate of Hawking emission (i.e. the “greybody factors”)? I can give no really satisfying answer to this question. Surely one has less control over the correspondence between strong and weak coupling away from the BPS sector. It is conceivable that the initial rates for Hawking radiation agree but the details about the correlations that develop over time do not match. In this scenario, there would simply be more non-BPS states at strong coupling than there are at weak coupling. This is not so hard to imagine, since in going from weak to strong coupling the causal structure of the background spacetime is distorted into that of a black hole. An analogy that may be useful is the coupling constant dependence of the state space of electrons in an atom. At sufficiently strong electric coupling the ground state becomes unstable and electrons can be absorbed into the nucleus, at which point the nuclear Hilbert space comes into play in resolving the physics. A strength of this analogy is that in the black hole case the ergoregion inside the (non-extremal) horizon also manifests a kind of instability of the ground state. #### AdS/CFT duality The near-horizon limit of the D-brane physics led to the celebrated Maldacena conjecture, according to which supergravity/string theory in an asymptotically Anti-deSitter spacetime is equivalent to a superconformal field theory on the conformal boundary of that spacetimemagoo . An example of this is the duality between superstring theory on $`AdS_5\times S^5`$ and a $`U(N)`$ super-Yang-Mills theory on $`S^3\times R`$, where $`N`$ is related to the string coupling $`g_s`$ the string length $`\mathrm{}_s`$ and the AdS radius $`R`$ by $`R^4=4\pi g_sN\mathrm{}_s^4`$, and the Yang-Mills and string couplings are related by $`g_{\mathrm{YM}}^2=4\pi g_s`$. There is much remarkable evidence in favor of the AdS/CFT duality, and no evidence against it to date. Hence, for the sake of argument, let us suppose it is valid and ask about the consequences for black holes. In the $`AdS_5\times S^5`$ example it has been shown that the entropy of a black hole which is large compared to the AdS radius (and is hence stable) is 3/4 of the entropy of a thermal state in the Yang-Mills theory at weak ’t Hooft coupling ($`g_{\mathrm{YM}}^2N1`$) at the corresponding Hawking temperature. Moreover, there is reason to believe that the entropy would only change by a factor of order unity if the calculation could be done at strong ’t Hooft coupling (which is what is required by for the case of large AdS radius). We thus have a puzzle similar to that in the case of the D-brane state counting, but now far from extremality. In the Yang-Mills theory it seems there can be no missing states corresponding to the degrees of freedom inside of the black hole. The entropy of the thermal state simply counts all states so, if the Maldacena conjecture is really true, one infers that there can be no independent degrees of freedom inside the black hole. Can this conclusion be evaded? A simple evasion is to suppose that the equivalence conjectured by Maldacena actually relates the supergravity observables only outside the horizon to the Yang-Mills observables in the boundary theory (see Fig. 1(a)). This would be consistent with causality and would certainly explain why all the states inside the black hole are not seen in the Yang-Mills theory. In fact, something like this seems almost necessary in view of the fact that the full Schwarzschild-AdS spacetime has a boundary with two disconnected pieces, the dark vertical lines in Fig. 1(a). What would be the role for the states in the Yang-Mills theory on the left if the one on the right already covered all the states inside the black hole? A different evasion is required if we consider not an eternal black hole but rather a black hole, small compared with the AdS radius, that forms from collapse and then evaporates (see Fig. 1(b)). In this case the AdS/CFT duality presumably states that all observables in the spacetime have correspondents in the CFT on the single boundary component. In particular, observables in the algebra $`C`$ localized behind the horizon (which is of course only defined relative to a particular state $`|\psi `$ of the CFT which corresponds to matter collapsing to form a black hole) must be contained within the full algebra $`B`$ of observables in the CFT. This in itself is not mysterious, since the field equations allow us to express any observable in $`C`$ as an observable in the algebra $`A_1`$ localized at a spacelike slice before the black hole ever formed. The question of whether there are independent states of the black hole interior is perhaps most sharply formulated here as the question whether the black hole evaporation is unitary from the viewpoint of the exteriorcftunitarity . Since the CFT itself is unitary, the question amounts to whether the algebra $`A_2`$ of observables on a spacelike slice after the black hole has evaporated completely is equal to $`A_1`$ or is rather a proper subalgebra of $`A_1`$. In the latter case one would need also the observables in the algebra $`C`$ behind the horizon to fill out the complete algebra. Moreover, causality would suggest that $`C`$ and $`A_2`$ would commute relative to the state $`|\psi `$, that is, expectation values of the ideal generated by the commutator algebra $`[C,A_2]`$ would vanish. Many practitioners of duality have argued that the equality $`A_2=A_1`$ is assured because the situations before collapse and after evaporation are similar: Anti-de Sitter spacetime with some matter. In particular, the “initial” configuration could have been the result of a prior black hole formation and evaporation, or of many cycles of formation and evaporation. If each black hole has internal observables not captured on the outside after the black hole is gone, then one seems to be requiring that the CFT contains within it (relative for an appropriate state) commuting subalgebras of observables corresponding to an infinite number of such black hole interiors, all of which commute with an algebra of outside observables such as $`A_2`$. This requirement seems difficult to reconcile with reasonable expectations about the number of states in the CFT at a given energy. Can one really arrange a sequence of black hole formations and evaporations where each black hole is made from the Hawking radiation into which the previous black hole evaporated? If not, then a state which produces many black holes must contain to begin with energy corresponding to each black hole. In this case there there are perhaps more states so let us suppose, to be difficult, that one can indeed repeatedly refocus the Hawking radiation to form an endless cycle of black holes with a finite amount of energy. In this case either the collection of commuting interior subalgebras exists, or one must deny the independence of the interior observables. Most string theorists support the second alternative. I prefer the first since it requires only nonintuitive behavior of the unfamiliar strongly coupled, (astronomically) large N gauge theory, rather than gross violations of locality where they would not otherwise be expected. #### Matrix Theory Matrix theory (a candidate for a nonperturbative formulation of string theory) can purportedly describe formation and evaporation of black holes, and the theory is manifestly unitary. There seems to be no room in matrix theory for any states corresponding to the interior of a black hole, left over after all particles in the Hawking radiation have dissipatedlenny . I have not yet learned enough about matrix theory to think carefully about whether or not there is any loophole through which this conclusion can be evaded. ## Nature of the Bekenstein Bound The “Bekenstein bound”bekbnd on the entropy that can be associated with a closed 2-surface $`\mathrm{\Sigma }`$ is $$S_\mathrm{\Sigma }\frac{A_\mathrm{\Sigma }}{4\mathrm{}G}.$$ (1) This is (presently) a heuristic notion motivated by the generalized second law of thermodynamics as follows. Suppose that by tossing in a suitable arrangement of matter the surface $`\mathrm{\Sigma }`$ could be made to coincide with a slice of the horizon of a black hole. Then the entropy of that black hole would be $`A_\mathrm{\Sigma }/4\mathrm{}G`$, which would violate the second law unless the entropy $`S_\mathrm{\Sigma }`$ associated with $`\mathrm{\Sigma }`$ if the extra matter is not tossed in is less, i.e. unless the bound (1) holds. In describing the Bekenstein bound I was careful to refer to $`S_\mathrm{\Sigma }`$ as the entropy associated with $`\mathrm{\Sigma }`$, rather than the entropy contained within $`\mathrm{\Sigma }`$, since the meaning of the bound (1) inferred by the black hole formation argument depends on the interpretation of the black hole entropy. If the black hole entropy is the logarithm of the number of states of the the black hole including the interior states, then we infer a “volume bound” on the entropy contained within $`\mathrm{\Sigma }`$. If however, as argued above, the black hole entropy reflects only those states that can influence the exterior, then we infer only a “surface bound” on the surface states of $`\mathrm{\Sigma }`$. I do not consider the volume bound interpretation to be viable. Not only can it not be inferred from the second law with the surface interpretation of black hole entropy, but it seems contradicted by the example used in the first section of this paper: since the volume of the region interior to the surface could be arbitrarily large it could contain an arbitrarily large amount of entropy. It also suffers from a species problem, that is, the entropy inside could be arbitrarily large if the number of independent fields in nature is arbitrarily large (but see bekdowe for another point of view). (On the other hand, if the number of species is sufficient for an order unity violation of the bound, then a black hole would be unstable to explosive evaporation on a timescale of order the light crossing time, and so the original rationale for the bound would be losttaj2b .) As an important side remark, note that the black hole formation argument suggesting the bound (1) does not apply to every closed 2-surface, since not every such surface can be made to coincide with a slice of the horizon of a black hole. Consider for instance an outer trapped surface inside a black hole. The future pointing null congruences orthogonal to this surface are converging on both sides, whereas the horizon generators are always non-converging according to the area theorem. For another example, consider the intersection of the past light cones of two spacelike related points $`p`$ and $`q`$. The future pointing null congruences orthogonal to this intersection surface are converging (to $`p`$ and $`q`$) on both sides. (This surface is not compact, but one can build a compact 2-surface out of pieces like this.) The restriction on surfaces is certainly necessary for the volume interpretation of the bound (although as discussed above I do not consider this interpretation to be viable in any case), since otherwise it is easy to find surfaces with arbitrarily little area enclosing a large volume. For example, a trapped surface near the singularity of a Schwarzschild black hole can have arbitrarily small area and still bound a finite volume. For another example, one can make a spacelike surface of arbitrarily small area enclose any volume by wiggling the surface “up and down” in the timelike direction. An interpretation of the bound (1) that is neither a volume nor a surface interpretation has been proposed by Boussoraf . In this interpretation, $`S_\mathrm{\Sigma }`$ is the entropy crossing any segment of a null hypersurface, meeting $`\mathrm{\Sigma }`$ orthogonally, that is expanding towards $`\mathrm{\Sigma }`$. The validity of this bound in a variety of contexts has been argued for in Ref. raf . The volume bound interpretation of (1) suggests the “holographic principle”thooftbnd ; holoprin according to which all the physics in the volume should be describable by a theory on the bounding surface $`\mathrm{\Sigma }`$. The surface bound interpretation on the other hand does not have any holographic connotation. Bousso suggests that his bound motivates a holographic priciple which refers to the null surface segments, but these segments do not in general span the volume. It thus seems to me that the holographic principle, while it may be a property of quantum gravity and/or of the AdS/CFT duality, is not logically suggested by the Bekenstein bound. ## Black hole entropy without black holes I have argued above that black hole entropy is not determined by the number of internal states of the black hole, but rather by the number of states, associated with the presence of the horizon, that can influence the outside world. This suggests that the notion of black hole entropy should apply not just to black holes but to any causal horizon. In fact, some approaches to computing the entropy associated with horizons do yield the result $`1/4`$ per Planck area of a Rindler horizon or a deSitter horizon, both of which are observer dependent horizons. For example, in a recent paper Carlip carlipcft finds this result from the representation theory of a conformal subgroup of the diffeomorpism group associated with any (non-degenerate) Killing horizon, and he points out that the Euclidean path integral approach also yields an entropy for deSitter horizonsgibbhawk . Also, the black hole pair creation probability is weighted by $`\mathrm{exp}(\mathrm{\Delta }A_{\mathrm{accel}}/4)`$ where $`\mathrm{\Delta }A_{\mathrm{accel}}`$ is the associated increase of the area of an acceleration horizonhawkhoroross . This strongly suggests a state-counting role for the entropy of acceleration horizons, an idea which is further supported by calculations relating transition amplitudes for particle creation processes to the associated change of horizon areamasspare . (Ref. hawkhoro argues that one should not attribute an entropy to the acceleration horizon because of its observer-dependent nature. For the reason articulated in the concluding remarks, I do not subscribe to this viewpoint.) As a more direct way to establish the validity of horizon entropy without black holes, I will will now argue that there are general laws of horizon thermodynamics, strictly analogous to those for black holes, for a class of causal horizons which I will call “partial event horizons”. Recall that the global event horizon of an asymptotically flat spacetime is the boundary of the past of future null infinity $`^+`$. I define similarly a partial event horizon (PEH) as the boundary $`I^{}[p]`$ of the past of a single point<sup>3</sup><sup>3</sup>3One could of course consider the boundary of the past of any subset of $`^+`$. $`p^+`$. In flat spacetime a PEH is just a Rindler (acceleration) horizon, and in an asymptotically flat spacetime a PEH asymptotically approaches a Rindler horizon. Although a PEH has cross sections with infinite area, it satisfies Hawking’s classical area theorem in the local sense that the expansion of its null generators is nowhere negative. The proof is similar to but slightly simpler than that for the event horizon since the assumption of cosmic censorship can be applied directly to rule out the possibility that a null generator leaves the PEH before reaching $`^+`$. Thus changes in the area are nonnegative, so a PEH satisfies a classical “second law of horizon mechanics”. A quasistationary region of a PEH also satisfies a “first law of horizon mechanics” that is strictly analogous to the first law of black hole mechanics $`dM=(\kappa /8\pi )dA`$. This law for black holes can be understood in a quasi-local fashion, called the “physical process version” in Ref.waldredbook , which applies to variations away from a quasi-stationary configuration with approximate horizon generating Killing field $`\chi ^a`$. In this setting $`dM`$ is interpreted as the flux $`T_{ab}\chi ^a𝑑\mathrm{\Sigma }^b`$ of “boost energy” across the horizon or a part thereof. A generic PEH will possess many quasistationary regions, to which the physical process version of first law will apply for the same reason as for black hole horizons. (The normalization ambiguity of the boost Killing field scales both $`dM`$ and $`\kappa `$ in the same way, so the first law is independent of this ambiguitytos . Finally, as for the generalized second law, note that Sorkin’s proposal for the origin of the generalized second law described above applies to any causal horizon, and in particular it applies to a PEH. Moreover, it seems that all gedanken experiments supporting the generalized second law for quasistationary processes involving black hole horizons would apply as well to quasistationary regions of PEH’s. ## Concluding Remarks What distinguishes a black hole horizon from a more general causal horizon is that it is universally defined with reference only to the global causal structure of the spacetime. The absence of reference to particular observers or classes of observers is thus its key distinguishing feature. In practice, however, this universality is irrelevant. For example, the universe may be spatially compact, and yet we have no reservations in applying the laws of black hole thermodynamics to approximately isolated “black holes”. It is always we who divide the system into the “outside” and the “inside”. It thus seems entirely natural that the notion of black hole entropy extends to general causal horizons. This generalized notion of horizon entropy preserves the the formula $`S=A/4\mathrm{}G`$, whose universality is understood as arising from the ultraviolet dominance of the “density of surface states”, much as the universal form of the short distance limit of quantum field correlations is understood. ## Acknowledgements I am grateful to numerous colleagues for even more numerous discussions on the topics discussed here. This work was supported in part by the National Science Foundation under grants No. PHY98-00967 at the University of Maryland and PHY94-07194 at the Institute for Theoretical Physics.
no-problem/9908/cond-mat9908263.html
ar5iv
text
# Spectroscopy of nanoscopic semiconductor rings ## Abstract Making use of self-assembly techniques, we demonstrate the realization of nanoscopic semiconductor quantum rings in which the electronic states are in the true quantum limit. We employ two complementary spectroscopic techniques to investigate both the ground states and the excitations of these rings. Applying a magnetic field perpendicular to the plane of the rings, we find that when approximately one flux quantum threads the interior of each ring, a change in the ground state from angular momentum $`\mathrm{}=0`$ to $`\mathrm{}=1`$ takes place. This ground state transition is revealed both by a drastic modification of the excitation spectrum and by a change in the magnetic field dispersion of the single-electron charging energy. The fascination of ring-like atomic and quantum structures dates back to Kekule’s famous proposal of the structure of benzene. Particularly interesting are the magnetic properties of such non-simply-connected quantum systems, which are related to the possibility to trap magnetic flux in their interior. Trapping of a single flux quantum in a small molecule such as benzene is impossible with the magnetic fields available in today’s laboratories. In recent years, however, the availability of sub-micron solid-state ring structures has triggered a strong interest in the magnetic properties of rings, especially in view of the fact that even in the presence of scattering, the many-particle ground state becomes chiral in a magnetic field, which leads to so-called ”persistent currents”. The large body of theoretical work on the properties of quantum rings is accompanied by a number of ground breaking experimental investigations of the magnetic and transport properties of rings. These studies have been carried out in the mesoscopic range, where scattering still influences the phase coherent transport and a large number of quantum states are filled. To the best of our knowledge, no spectroscopic data is available on rings in the scatter-free, few electron quantum limit. Furthermore, despite a strong theoretical interest, the only data available on the excitations of rings were taken on macroscopic structures. Here, we report on the spectroscopy of the ground states and excitations of self-assembled, nanoscopic InGaAs quantum rings, occupied with one or two electrons each, and subjected to magnetic fields $`0B12`$ T, corresponding to 0 – 1.5 flux quanta threading the interior of the ring. In both ground state and excitation spectroscopy we observe characteristic changes at magnetic fields around $`B=8`$ T which are attributed to the development of a magnetic-field-induced chiral ground state. The quantum rings are fabricated by solid-source molecular-beam epitaxy, using the Stranski-Krastanov growth mode, which has now become a well-established technique for the fabrication of high-quality, self-assembled semiconductor nanostructures. Recently, we have reported on a remarkable change in morphology when InAs self-assembled dots, grown on GaAs, are covered with a thin layer of GaAs and annealed at growth temperature (1 min. at 520 C for the present samples) . Then, the shape of the dots drastically changes from that of a lens (roughly 20 nm in diameter, 7 nm in height) to one resembling a volcano, with an increased lateral size (between 60 and 140 nm in outer diameter), a reduced height (around 2 nm) and a well-defined center hole of around 20 nm diameter - see also inset in Fig. 1. For the present investigation, these ”self-assembled quantum rings” are embedded in a field-effect transistor (FET) structure, which makes it possible to control the number of electrons per ring by application of a suitable bias voltage. The electron number can be monitored by capacitance-voltage (CV) spectroscopy, which also allows us to derive the many-particle ground state energies of the $`n`$-electron system. Furthermore, far-infrared (FIR) transmission spectroscopy can be performed simultaneously, to obtain the excitation energies of the electronic system. It should be pointed out that even though the fabrication of self-assembled rings has been demonstrated before, no conclusive evidence was available to show that the surface ring morphology translates into an electronic ring structure inside the completed FET. The present FET layer sequence starts with a highly Si-doped GaAs back contact layer and a 25 nm GaAs spacer, followed by the InAs Stranski-Krastanov layer, a 30 nm GaAs cover layer, a 116 nm thick AlAs/GaAs superlattice, and a 4 nm GaAs cap. Details of the layer sequence and growth procedure, which are essentially identical to those used for quantum dots, can be found in Refs. . The crucial step for the ring formation is an interruption in the growth of the GaAs cover layer after the deposition of a nominal thickness $`\theta `$. On reference samples, which at this stage were removed from the growth chamber, we observe ring formation for $`\theta `$ between 1 and 4 nm. The inset in Fig. 1 shows an atomic force micrograph of a reference sample with $`\theta =2`$ nm. The electronic spectroscopy was performed at liquid He temperatures on samples with an effective area of around 5 mm<sup>2</sup>, covering approximately $`5\times 10^8`$ rings. The CV traces were taken at low frequencies ($`<1`$ kHz) using standard lock-in technique; the FIR response was recorded by a rapid-scan Fourier transform spectrometer. The main part of Fig. 1 displays CV spectra of samples with different $`\theta `$ but identical layer sequence. These spectra demonstrate the strong influence that $`\theta `$ has, not only on the morphology of the bare rings, but also on the electronic properties inside the completed FET structure. For $`\theta `$ = 5 nm, the spectra are indistinguishable from those of common quantum dot samples without growth interruption: We observe a double peak structure at a gate voltage $`V_g1.1`$ V, corresponding to the single electron charging of the two spin states of the so-called ”$`s`$-shell” in the dots. Because of inhomogeneous broadening in the present, large-area samples, the 4 maxima of the $`p`$-shell cannot be well distinguished and have merged into a broad plateau around $`V_g0.4`$ V. For details on quantum dot spectroscopy, see, e.g., Refs. . For the sample with $`\theta `$ = 1 nm, two peaks can be clearly distinguished around $`V_g0`$ V. For $`\theta =3`$ nm, no individual peaks can be identified and only one very broad structure is observed around $`V_g0.5`$ V. We attribute this to the fact that at the threshold of ring formation, the morphology of the Stranski-Krastanov layer is not well defined. Nevertheless, this data is of significance, as it is a further indication for the structural change that takes place when the growth is interrupted. The shift of the first maximum from $`1.2`$ V ($`\theta =5`$ nm) to around 0 V ($`\theta =1`$ nm) can be explained by an upward shift of the ground state energy caused by the reduced height of the rings compared to the dots. In the following, the discussion will focus on the $`\theta =1`$ nm ring sample. Figure 2(a) shows the normalized FIR transmission of this sample at $`V_g=0.143`$ V (upward arrow in Fig. 1) for two different magnetic fields $`B`$, applied perpendicular to the plane of the rings. Comparing the carrier density obtained from either CV or FIR spectroscopy with the ring density determined by AFM, we find that at this gate voltage, each ring is filled with approximately $`n_e=2`$ electrons, which shows that (as for the dots) each CV maximum corresponds to the filling of one electron per ring. It should be pointed out here that the FIR measurements, which require a signal-to-noise ratio of around 1 part in $`10^4`$, are extremely challenging and at the limit of state-of-the-art FIR spectroscopy. However, from a thorough evaluation of a large number of spectra, including subtraction of the superimposed signal from the cyclotron resonance in the back contact, we can obtain the FIR response as a function of the magnetic field, which is shown in Fig. 2(b). As indicated by the different symbols, the resonances in Fig. 2(b) can be grouped into the following modes: two resonances ($``$) which are degenerate at $`B=0`$ and exhibit orbital Zeeman splitting when a magnetic field is applied. A low-lying mode ($``$) which, due to an insufficient signal-to-noise ratio at very low energies, can only be detected above 10 meV, but extrapolates to $`7`$ meV at $`B=0`$. This mode dies out around $`B=7`$ T, when also the lower $``$-mode vanishes and a new mode ($`\mathrm{}`$) appears. The resonances summarized in Fig. 2(b) differ quite strongly from those observed in quantum dots, where in general, only two resonances are observed, one of which increases with increasing field, whereas the other decreases. On the other hand, Fig. 2(b) can be directly compared to the excitation spectrum of quantum rings, as calculated, e.g., by Halonen et al.. Even though these calculations were performed for a ring with much larger dimensions, all the above experimental features are in good qualitative agreement with the calculated energy dispersion. Furthermore, using the effective mass of $`m^{}0.07`$ $`m_e`$ found in other self-assembled In(Ga)As quantum structures, we find that the slopes of the $``$-modes ($`\pm \frac{1}{2}\mathrm{}\omega _c`$, solid lines), the $``$-mode ($`\frac{1}{2}\mathrm{}\omega _c`$, dotted line), and the $`\mathrm{}`$-mode ($`\mathrm{}\omega _c`$, dashed line) are all in agreement with the calculations. From this we conclude that indeed the morphology seen in Fig. 1 is preserved when the growth is continued, and that it translates into a ring-like electronic structure. As will be discussed in the following, the change in the spectrum around $`B=8`$ T can then be understood as a direct consequence of a magnetic-field-induced change in the ground state. The $`\times `$-mode is not found in the calculated ring excitations of Ref. . At present, an explanation for these resonances is still missing, and it cannot be ruled out that these resonances originate from the presence of a few large quantum dots which have not developed into rings and therefore do not change their ground state at $`B=8`$ T. The electronic states in rings can be discussed using a simple model of a circular, 1-dimensional wire, bent into a circle of radius $`R`$. The energy levels then follow from the periodic boundary conditions to $`E_{\mathrm{}}=\frac{\mathrm{}^2}{2m^{}}k_{\mathrm{}}^2`$ with $`k_{\mathrm{}}=\mathrm{}\frac{1}{R}`$. When a flux $`\varphi =\pi R^2B`$ penetrates the interior of the ring, an additional phase is picked up by the electron on its way around the ring, which leads to $$E_{\mathrm{}}=\frac{\mathrm{}^2}{2m^{}R^2}(\mathrm{}+\frac{\varphi }{\varphi _0})^2,\mathrm{}=0,\pm 1,\pm 2,\mathrm{}$$ (1) with $`\varphi _0`$ being the flux quantum. Thus, with increasing magnetic field, the ground state will change from angular momentum $`\mathrm{}=0`$ to one with higher and higher negative $`\mathrm{}`$ (see Fig. 3(a)), a fact intimately related to the persistent currents in mesoscopic rings. Furthermore, a periodic, Aharonov-Bohm-type oscillation in the ground state energy will take place, as shown in Fig. 3(a). Obviously, each change in the ground state will in turn lead to a pronounced change in the possible transitions, such as the one seen in Fig. 2(b) at $`B8`$ T. For a more quantitative description of the energies and transitions in the present rings, we have used a model by Chakraborty et al. and calculated the single-particle states in a ring potential $`U(r)=\frac{1}{2}m^{}\omega _0^2(rR_0)^2`$ (see inset in Fig. 3(b)), where $`\omega _0`$ is the characteristic frequency of the radial confinement and $`R_0`$ is the radius of the ring. These two parameters determine the low energy resonance at $`B=0`$, estimated at around 7 meV and corresponding to an azimuthal excitation $`\mathrm{\Delta }\mathrm{}=\pm 1,\mathrm{\Delta }N=0`$, as well as the high energy resonance at 20 meV, corresponding to an excitation $`\mathrm{\Delta }\mathrm{}=\pm 1,\mathrm{\Delta }N=1`$, where $`N`$ is the radial quantum number. In this way, both parameters are readily determined to be $`R_0=14`$ nm and $`\mathrm{}\omega _0=12`$ meV. The electronic radius of 14 nm is in good agreement with the effective radius of the uncovered rings (i.e. the radius where the InGaAs has maximum thickness, see Fig. 1), which is roughly 18 nm. Furthermore, we find satisfactory agreement between the calculated and measured transitions over the entire range of magnetic fields. Using more realistic potentials would certainly improve this agreement, however at the cost of introducing additional adjustable parameters. The lines in Fig. 3(b) represent the magnetic-field dispersion of the ring energy levels, calculated using the model potential $`U(r)`$ and the values for $`R_0`$ and $`\omega _0`$ derived above. It can be seen that the calculations predict a ground state transition (arrow) from $`\mathrm{}=0`$ to $`\mathrm{}=1`$ at $`B8`$ T, in good agreement with the position of the change in the FIR resonances. This agreement is somewhat surprising, considering that the FIR data was obtained for $`n_e=2`$, whereas the model is single-particle. From this we conclude that (as in self-assembled InAs dots) the single particle states are a quite accurate basis for the description of the many-particle states and excitations. This assumption is supported by the fact that the measured FIR resonance positions for $`n_e=1`$ are very similar to those for $`n_e=2`$, even though the signal-to-noise ratio for $`n_e=1`$ does not allow us to clearly identify all resonances shown in Fig. 2(b). There is an additional way to confirm the ground state transition associated with the trapping of flux inside the rings. By carefully evaluating the position of the lowest capacitance maximum (downward arrow in Fig. 1), we have a direct experimental access to the $`n_e=1`$ ground state energy. The data points in Fig. 3(b) summarize the shift in gate voltage of the lowest capacitance peak as a function of the magnetic field. As indicated by the arrow, at $`B=8.2`$ T, corresponding to a flux of $`\pi R_0^2B\varphi _0`$, a change in the slope can be identified in the data. From the comparison with the model calculations, this change in slope is identified as the magnetic-field-induced ground state transition from $`\mathrm{}=0`$ to $`\mathrm{}=1`$. Note that the shift of the $`n_e=1`$ charging peak is given in mV (right hand scale in Fig. 3(b)). From a comparison with the calculations (left hand scale) we find a voltage-to-energy conversion factor of $`f=e\mathrm{\Delta }V_g/\mathrm{\Delta }E=1.8`$. Using the lever arm model, which has proved to be quite accurate in the case of quantum dots , the present layer structure gives $`f=7`$. We attribute this discrepancy to the fact that the slopes at which the $`\mathrm{}=0`$ and $`\mathrm{}=1`$ states intersect strongly depends on the detailed choice of the confining potential (as can readily be seen from a comparison between Figs. 3(a) and (b)). This sensitivity could in fact provide for a useful handle to derive the confining potential from more elaborate model calculations. Converting the separation between the lowest charging peaks in Fig. 1 into energy, we obtain a Coulomb interaction energy of roughly 20 meV for both rings and dots. This similarity is somewhat surprising, given the larger lateral size of the rings. The missing central part, which will decrease the effective area, may partly be responsible for the large Coulomb interaction in the rings. Here, more in-depth theoretical work is desirable, especially in view of a possible formation of a rotating Wigner molecule and with respect to the fact that experimentally the 1- and 2-electron excitation spectra are very similar. In summary, we have used the striking morphological change that takes place when self-assembled InAs quantum dots are partially covered with GaAs to fabricate nanoscopic quantum rings with dimensions that bridge the size range between mesoscopic and molecular ring structures. The electron states in these rings are not effected by the presence of random scatterers and are dominated by quantum effects rather than Coulomb interaction. Using two complementary spectroscopic techniques, we have identified a magnetic-field-induced transition from a ground state with zero angular momentum to a chiral ground state. This transition is a direct consequence of the non-simply-connected ring geometry and takes place when approximately 1 flux quantum penetrates the effective interior area of the rings. We would like to thank R. J. Warburton, K. Karraï, T. Chakraborty, V. Gudmundsson, and M. Barranco for helpful and inspiring comments, the latter also for making their unpublished work available to us. Financial support from QUEST, a NSF Science and Technology Center, and from BMBF, through Grant 01 BM 623 and a Max Planck research award, is gratefully acknowledged.
no-problem/9908/astro-ph9908263.html
ar5iv
text
# Ab Initio Formation of Galaxies, Groups and Large-Scale Structure11footnote 1Invited talk at the IAU Colloquium 174 “Small Galaxy Groups”, Turku, Finland, 13th-18th June 1999. ## 1 Introduction Groups of galaxies probe the intermediate scale between galaxies and clusters. A satisfactory theory for the formation and the evolution of these systems has always been difficult, because the size of groups is small enough to require the modelling of the properties and the internal structure of the galaxy members and, at the same time, large enough to require the modelling of the large scale structure sorrounding the groups. For the last twenty years, $`N`$-body simulations have modelled groups with vacuum boundary conditions in order to resolve the internal kinematics of the galaxies (e.g. Carnevali et al. 1981; Barnes 1985; Diaferio et al. 1993; Athanassoula et al. 1997). Large-scale structure simulations have provided cosmological boundary conditions without resolving the galaxy internal properties (e.g. Frederic 1995a,b; Nolthenius et al. 1997). Both sets of simulations have not modelled the physics of galaxy formation and evolution which provides observable quantities, such as luminosity and colour. However, modelling groups within a cosmological context with sufficient resolution is extremely important, because groups both trace the large-scale structure of the Universe (e.g. Ramella et al. 1997, 1999) and are the sites of galaxy interactions (Hickson 1997; but see also e.g. Bettoni; Bosma; Rampazzo; Temporin, these proceedings). Galaxy interactions are connected with the formation of AGN’s (Byrd, these proceedings), quasars (Cavaliere, these proceedings) and multiple central black holes (Valtonen, these proceedings). The importance of this large–small scale connection has been particularly important for successfully modelling the very existence of compact groups (Mamon 1986, 1989; Diaferio et al. 1994; Hernquist et al. 1995). The enormous progress of the last few years in probing the high-redshift Universe, which shows a high degree of galaxy clustering (see e.g. Mazure & Le Fèvre 1999), impels us to build a cosmologically motivated theory of galaxy formation. State-of-the-art $`N`$-body/hydrodynamic simulations have started only recently to barely reach the resolution required to follow the formation of galaxies within a cosmological context, even though only some of the relevant physical processes are included (Pearce et al. 1999). For the time being, a more fruitful approach has been the combination of semi-analytic modelling of galaxy formation with $`N`$-body simulations. The semi-analytic approach enables us to have full control of the relevant galaxy formation processes, which occur on parsec or smaller scales, namely star formation, stellar evolution, and effects from supernova explosions. The $`N`$-body simulations yield the merging history of galaxies and their phase space position, which are the result of group and cluster formation occurring on megaparsec and larger scales. By combining the two techniques, we are able to connect the stellar population properties of galaxies with their clustering and kinematic properties. This approach has already been particularly successful (Kauffmann et al. 1999a,b; Diaferio et al. 1999). Here, I briefly review this technique (Section 2). I then show some remarkable results which the self-consistency of the approach yields automatically: the two component luminosity functions of groups (Section 3) and the density and velocity biases of cluster galaxies with differing colours (Section 4). Extracting mock redshift surveys from the simulation box also allows a direct comparison with the observed large-scale distribution of galaxies (Section 5). These results show the effectiveness of our approach, and also the partial success of the cosmological models. ## 2 The Galaxy Formation Recipe According to standard inflationary models (e.g. Peacock 1999), large scale structure in the present day Universe originates from primordial density perturbations amplified by gravitational instability. These models also predict that most of the mass in the Universe is non-baryonic and therefore “dark”. Assuming a Cold Dark Matter (CDM) power spectrum of the perturbations (e.g. Efstathiou et al. 1992), the clustering proceeds hierarchically: large perturbations collapse first, forming small dark matter halos which aggregate to form larger halos. Within these halos the baryonic gas cools, flows into the centre of the halo, and forms stars which evolve and, eventually, some of them explode as supernovae. The supernovae enrich the interstellar gas with metals and reheat part of the gas. We need two basic ingredients to implement this picture: (1) the merging history of dark matter halos; and (2) recipes for the relevant processes involving the baryonic matter. Traditional semi-analytic models (e.g. Somerville & Primack 1999 and references therein) use Monte Carlo simulations based on the extended Press & Schechter (1974) formalism (e.g. Bower 1991) to derive the merging tree of dark matter halos. Here, instead, we derive the merger trees from $`N`$-body simulations. Moreover, we identify the galaxy harbored by the halo with the central dark matter particle of the halo: the galaxy and the central particle share the same phase space coordinates. When two halos merge, the galaxy of the most massive halo “jumps” onto the central particle of the resulting halo. The other galaxy still retains its identity with its original dark matter particle, and becomes a “satellite” within the new halo. This approach makes the trajectory of galaxies in phase space discontinous, but it assures a smooth evolution of the luminosity and stellar mass of the galaxies. We then implement the semi-analytic technique as follows. We assume that the gas is in hydrostatic equilibrium within the dark matter halo. We model both the baryonic and the dark matter components with truncated isothermal spheres. The gas cools and flows on to the central particle of the halo instantaneously, if the cooling time is shorter than the Hubble time at that epoch, or at a given rate, otherwise. The stellar mass $`M_{}`$ increases accordingly to $$\frac{dM_{}}{dt}=\alpha M_{\mathrm{cold}}\times 10\frac{V_c}{R_{\mathrm{vir}}},$$ (1) where $`M_{\mathrm{cold}}`$ is the mass of cold gas, $`R_{\mathrm{vir}}`$ is the radius of the halo of mass $`M_{\mathrm{vir}}`$, and $`V_c^2=GM_{\mathrm{vir}}/R_{\mathrm{vir}}`$ is the halo circular velocity. The star formation efficiency $`\alpha `$ is a free parameter. Given the initial mass function for the stellar population (Scalo 1986), we can compute the number $`\eta _{SN}`$ of supernovae of Type II expected per unit mass. We neglect supernovae of Type Ia. A fraction $`ϵ`$, kept as a free parameter, of the kinetic energy $`E_{SN}`$ ejected by each supernova reheats part of the cold gas to the halo virial temperature. The mass of the reheated gas over the time $`\mathrm{\Delta }t`$ is $$\mathrm{\Delta }M_{\mathrm{reheat}}=ϵ\frac{dM_{}}{dt}\frac{4\eta _{SN}E_{SN}}{3V_c^2}\mathrm{\Delta }t.$$ (2) Satellites merge with the central galaxy on a timescale set by dynamical friction. We compute this timescale analytically. Our prescription does not allow for merging between satellites. This shortcoming is partly responsible for the overproduction of central bright galaxies, along with the fact that we do not implement a model of chemical evolution (e.g. Kauffmann & Charlot 1998), which affects the gas cooling rate (see Section 3). Luminosity evolution in different bands is computed with stellar population synthesis models (Bruzual & Charlot 1999). Finally, galaxy luminosities are dimmed with an empirical prescription for dust extinction (Cardelli et al. 1989). We apply our technique to two variants of a CDM Universe: a $`\tau `$CDM and a $`\mathrm{\Lambda }`$CDM Universe. A complete description of these models and our technique is in Kauffmann et al. (1999a). ## 3 The Luminosity Function of Galaxy Systems A scientific theory is required to provide testable predictions. Sometimes, it also automatically provides explanations for phenomena for which the theory was not originally constructed. For example, consider the luminosity function of galaxies within different environments. It is well known that the usual Schechter function (Schechter 1976) does not fit all galaxy samples well. Many clusters (e.g. Biviano et al. 1995; Molinari et al. 1998; Trentham 1998), rich groups (Koranyi et al. 1998) and even compact groups (Hunsberger et al. 1998; see, however, e.g. de Carvalho et al. 1994 and Zepf et al. 1997) show a two component luminosity function: a roughly Gaussian distribution at the bright end, superimposed on a Schechter function which dominates at the faint end. Because ellipticals and S0’s are generally more luminous than spirals and irregulars, the Gaussian component is mostly populated by early type galaxies, and the Schechter component by late type galaxies. Moreover, because this bump at the bright end peaks at different magnitudes depending on the system considered, when we consider a large galaxy sample, which includes different environments, this feature disappears, and a Schechter function provides a reasonably good fit (e.g. Marzke et al. 1994). Simulations of the formation of groups and clusters, where galaxies are resolved and are allowed to merge, predict a typical two component mass function at later times. This result has been well known since very early $`N`$-body simulations (Aarseth & Fall 1980; Roos & Aarseth 1982: Cavaliere et al. 1991). Thus, on the assumption that light is proportional to mass, galaxy merging provides a natural explanation for the bump observed in the luminosity function of galaxy systems. The fact that this bump is not always present clearly contains information on the formation history of the galaxy system. Our models, designed to reproduce the global properties of galaxies and their large scale distribution rather than the luminosity function of individual clusters, also show this characteristic bump for massive systems (long and short dashed lines in Fig. 1). Apparently, the global luminosity function (bold solid line) shows no bump. Note that the global luminosity function is not a Schechter function. To obtain the Schechter form, it is sufficient to use a dynamical friction prescription derived from high resolution $`N`$-body simulations (Springel et al. 1999), which automatically includes merging between satellites (Somerville & Primack 1999), rather than our analytic prescription (Section 2). The agreement with the Schechter function can be improved further by modelling the chemical enrichment of the interstellar medium, which reduces the cooling efficiency at early times. ## 4 Morphology-Density Relation and Velocity Bias The morphology-density relation has been known since the very beginning of the systematic investigation of galaxies (e.g. Hubble & Humason 1931). It has also been known that late type galaxies within clusters have larger velocity dispersion than early type galaxies (e.g. Moss & Dickens 1977). Modelling galaxy and large scale structure formation at the same time provides a self-consistent way of explaining these properties of cluster galaxies. Fig. 2 compares our models with the surface number density profiles of galaxy clusters. Profiles are for blue and red galaxies (CNOC sample, Carlberg et al. 1997) and galaxies with emission and non-emission line spectra (ENACS, Biviano et al. 1997). The agreement is satisfactory. In fact, the differences between models and observations are simply due to the galaxy subsample definition: real galaxies have been selected according to their spectral properties which we do not model here. The models can also account for the larger velocity dispersion of blue galaxies compared to red galaxies. Blue galaxies are falling on to the cluster for the first time and will become red later, when they have exhausted their cold gas reservoirs. Although this plausible explanation is well-known (e.g. Mohr et al. 1996), our models yield this result self-consistently. Moreover, our models show how this result depends on the cosmological model (Diaferio et al. 1999, Diaferio 1999). Our technique is therefore particularly promising. For example, when simulations with better spatial and force resolution become available, it will be possible to investigate those kinematic properties of compact group members which, at present, seem difficult to interpret. For example, high velocity galaxies (Sulentic, these proceedings) may simply be due to favourable, albeit rare, conditions of infall from the surroundings (Diaferio et al. 1994). ## 5 Mock Redshift Surveys Our simulation boxes have volumes that are comparable to the volume probed by the northern region of the Center for Astrophysics magnitude limited Redshift Survey (CfA2N hereafter; Geller & Huchra 1989; Huchra et al. 1990; de Lapparent et al. 1991; Huchra et al. 1995; Falco et al. 1999). Therefore, we are able to compare our models with a large portion of the Universe by analysing our simulation box the same way observers analyse their data. We have compiled mock redshift surveys in order to have a cluster as massive as Coma at the same location as in the CfA2N (see Diaferio et al. 1999 for details). Fig. 3 shows the CfA2N and Fig. 4 shows a mock catalogue from one of our simulation boxes. The large scale structure in our simulation is not as sharply defined as in the real Universe: for example no Great Wall (Geller & Huchra 1989) is present in the model. Schmalzing & Diaferio (1999) have quantified these topological differences on large scale using Minkowski functionals. Further constraints on the model of galaxy and structure formation come from clustering on smaller scales, in particular from the properties of galaxy groups (Ramella et al. 1997, 1999; Tucker, these proceedings). Catalogues of groups are usually compiled using a friends-of-friends algorithm on redshift data (Huchra & Geller 1982). However, in these redshift surveys, we do not know the true distance to a galaxy; we only know its redshift, which also includes the galaxy peculiar velocity. Therefore, the average properties of groups extracted from a redshift survey do not necessarily agree with those of groups extracted from an ideal survey in configuration space. $`N`$-body simulations allow us to quantify this difference. Although previous analyses suggest that the differences are small (e.g. Nolthenius & White 1987; Moore et al. 1993; Frederic 1995a, 1995b; Nolthenius et al. 1997), these analyses do not include the physics of galaxy formation. On the other hand, we are able to compile redshift surveys that take into account the luminosities of galaxies, which result from their merging and stellar evolution histories. We find that typical parameters of the friends-of-friends algorithm yield kinematic properties which are in agreement with both those of groups selected from configuration space and those of the CfA2N groups (Fig. 5). However, 40% of the triplets selected in redshift space are, in fact, unrelated galaxies which are not bound systems in configuration space. Some of the differences between groups in the models and in the CfA2N is due to the differing distribution of galaxies on large scale. An $`N`$-body simulation constrained to have the same large scale structure as the local Universe (Mathis et al. 1999) will be valuable for further constraining the galaxy formation recipes using the properties of groups. ## 6 Conclusion Combining semi-analytic modelling with $`N`$-body simulations is an extremely powerful tool for investigating galaxy and group formation within a self-consistent cosmological framework. $`N`$-body simulations with better spatial and force resolution than the simulations presented here will be ideal for understanding the properties of compact groups, their different morphology and environment (e.g. de Carvalho, these proceedings), their abundance in the local Universe and at high redshift. Groups selected objectively from two-dimensional information (Prandoni et al. 1994), with follow-up redshift measurements (Iovino, these proceedings), or directly selected in redshift space (Barton et al. 1996; Allam, these proceedings), can provide a wealth of information on both cosmology and galaxy formation processes. The physics of the intergalactic medium should also be included in our approach (Cavaliere et al. 1998; Cavaliere, these proceedings): both HI (e.g. Combes; Verdes-Montenegro; Sancisi, these proceedings) and X-ray observations (e.g. Mulchaey; Ponman, these proceedings) can show clear signatures of galaxy interactions and tell us about the dynamical state of the group (Diaferio et al. 1995). To simulate the evolution of a group of galaxies, we can extract the initial conditions and the external tidal field from a large-scale $`N`$-body simulation. This strategy has already been extremely successful in simulating galaxy clusters (Tormen et al. 1997; Springel et al. 1999). Thus, high-resolution $`N`$-body/hydrodynamic simulations will enable us to predict the properties of the galaxies and the intergalactic medium of individual groups within a full cosmological context. ###### Acknowledgements. The $`N`$-body simulations were carried out at the Computer Center of the Max-Planck Society in Garching and at the EPPC in Edinburgh, as part of the Virgo Consortium project. I thank Jörg Colberg, Guinevere Kauffmann and Simon White for permission to reproduce results from our joint research programme within the German Israeli Foundation (GIF) collaboration. I thank Margaret Geller and John Huchra for allowing the comparison of the CfA2N data directly with the simulations, and Ravi Sheth for a careful reading of the manuscript. I thank Mauri Valtonen for the invitation to this meeting, and him and the other organizers for a stimulating conference and a delightful time in Turku. Special thanks to Reinaldo de Carvalho and Giorgio Palumbo for lovely discussions and unexpected lightening endings of some Finnish nights.
no-problem/9908/nucl-ex9908015.html
ar5iv
text
# EXPERIMENTAL AND COMPUTER SIMULATION STUDY OF RADIOACTIVITY OF MATERIALS IRRADIATED BY INTERMEDIATE ENERGY PROTONS ABSTRACT The results of measurements and computer simulations of radioactivities and dose rates as functions of decay time are presented for <sup>nat</sup>Pb and <sup>209</sup>Bi irradiated by 1.5-GeV protons, <sup>59</sup>Co, <sup>63</sup>Cu, and <sup>65</sup>Cu irradiated by 0.13- and 1.2-GeV protons, and <sup>232</sup>Th and <sup>nat</sup>U irradiated by 0.1- and 0.8-GeV protons. The activities and dose rates are measured by direct high-precision $`\gamma `$-spectrometry. The irradiations were made using external beams extracted from the ITEP U-10 proton synchrotron. Simulations made using the LCS and CINDER’90 code systems are compared with measurements. INTRODUCTION The design of a hybrid system driven via a high current accelerator needs to have information about residual nuclides that are produced by high energy protons in the target and in constructional elements, defining a number of technical features of the system. Our previous works present the results of experimental measurements of residual radioactive nuclides produced in the target and constructional materials irradiated with 0.13- and 1.5-GeV protons. Predictive powers of a number of high energy codes were investigated using the experimental data. A general conclusion found insufficient accuracy of the codes to simulated radionuclide yields, both independent and cumulative, with the mean deviation of simulated results from experimental data often of a factor of 2 or higher. Reliable information on the yield of each residual nuclide is very important and such characteristic can be defined as “individual”. Unlike our previous works, this study is aimed to investigate the predictive power of codes not for individual but for “integral” parameters – time dependent activation and dose-equivalent rate (called dose rate hereafter) which are directly dependent on summed yield of individual radionuclides. For such “integral” characteristics, one may expect a much better predictive power of codes than in simulations of “individual” parameters. We believe that the dose rate due to eminations of the $`\gamma `$-s from any hybrid system part being irradiated is one of the most important working parameters in applications. Here, the LAHET and HMCNP Monte Carlo transport codes were used together with the code CINDER’90 to simulate the build-up and decay of the products. EXPERIMENT The irradiations of experimental samples were carried out using external beams of the ITEP proton synchrotron. A sandwich of experimental and aluminum samples of the same diameters of 10.5 mm was placed perpendicularly to the proton beam during each irradiation. The parameters of irradiations and of experimental samples are presented in Table 1. The CANBERRA spectrometer based on the GC-2518 Ge detector was used to measure $`\gamma `$-spectra of irradiated experimental and aluminum samples. $`\gamma `$-spectra processing and nuclide identification were made using the ASPRO and SIGMA codes and the PCNUDAT nuclear database. <sup>24</sup>Na formation in aluminum samples was used to determine the irradiation time integrated proton flux. The details of irradiation techniques, schemes and parameters of external beams, techniques of nuclides identification and their cross section determination can be found in . SIMULATION TECHNIQUE The simulation of build-up during irradiations and following decay of produced nuclides was made using the LCS and CINDER’90 code systems. LCS includes the LAHET and HMCNP Monte-Carlo radiation transport codes. LAHET follows protons of all energies and secondary neutrons above 20 MeV throughout the geometry using a selection of on-line nuclear models and parameters; for these calculations LAHET2.83 was used with default selections except for use of ISABEL intranuclear-cascade model with default parameters, preequilibrium model following intranuclear cascade, and nuclear elastic scattering for neutrons and protons. LAHET passes the position, direction, and energy of all neutrons below 20 MeV as source particles to HMCNP, which uses evaluated cross-section data to follow all neutrons (and photons, if desired) throughout the geometry. The regional nuclide production (At./proton) from LAHET and the binned neutron fluence (n/bin-p-cm<sup>2</sup>) from HMCNP are scaled by region volume (cm<sup>3</sup>) and proton source strength (p/s) to form production probabilities (At./cm<sup>3</sup>-s) and neutron fluxes (n/cm<sup>2</sup>-s). These are used with evaluated cross sections and decay data in CINDER’90 to calculate the temporal nuclide inventory during and following irradiation. In these simulations of thin-target measurements in environment having no features contributing to lower-energy neutron flux, contributions to nuclide inventory from the lower energy neutron reactions in HMCNP and in CINDER’90, although included, are negligible. To investigate the real predictive power of the codes but not just an ability to describe some known measured data, all simulations have been performed at Los Alamos strictly before receiving any experimental data from Moscow that could be compared with theoretical results. PROCESSING EXPERIMENTAL AND SIMULATED DATA Determination of experimental activities and dose rates Activity of each nuclide in experimental samples was calculated via least square approximating nuclide counting rates of corresponding $`\gamma `$-energy peaks taking into account $`\gamma `$-yields and absolute effectiveness of the spectrometer at corresponding $`\gamma `$-energy. The total activities of the targets were defined as a sum of activities of individual nuclides. The individual activities were determined dependently of properties of precursors present for practically all measured products: 1) If precursors are absent or have lifetimes that are either much less than the first measurement time after irradiation or much higher than the last measurement time, the following formula was used for activity calculation: $$a(t,T)=\frac{A_0}{\epsilon \eta }\mathrm{exp}[\lambda (tT)]\text{ .}$$ 2) If precursors have lifetimes that are comparable with times passed after irradiation, the following formula was used: $$a(t,T)=\frac{1}{\epsilon \eta }\{A_1\mathrm{exp}[\lambda _1(tT)]+A_2\mathrm{exp}[\lambda _2(tT)]\}\text{ ,}$$ where, $`A_0`$, $`A_1`$, and $`A_2`$ are the coefficients determined via the least square approximation; $`\lambda `$, $`\lambda _1`$, and $`\lambda _2`$ are the decay constants of the nuclide and its precursors; $`t`$ is the time after starting the irradiation; $`T`$ is the irradiation duration; $`\eta `$ is the $`\gamma `$-yield; $`\epsilon `$ is the absolute effectiveness of the spectrometer. The dose rate at 1 cm distance from a target was defined via summing individual nuclides dose rates which were determined via individual activities multiplied by $`K_\gamma `$-dose coefficients calculated for each nuclide using methods and data described in . Simulated activities and dose rates To have a correct comparison of obtained experimental results with the simulated ones, the simulated total activities were calculated by summing the individual simulated activities of only the nuclides that were identified in measurements and used for calculating experimental activities. The simulated dose rates were calculated by summing the individual dose rates of the same nuclides. AGREEMENT BETWEEN SIMULATED AND EXPERIMENTAL RESULTS The determined experimental and simulated activities are presented in Figures 1–6. Additionally, the figures show the mean squared deviation factors calculated via $$F=10^{\sqrt{\left(\mathrm{lg}\left({\displaystyle \frac{\text{theor. value}}{\text{exp. value}}}\right)\right)^2}}$$ for each of the data sets compared, where $`\mathrm{}`$ designates averaging over all of the comparison events. As seen from the figures, the simulated results are in good agreement with experimental data for most of the targets. In most cases, the mean squared deviation factor is from 1.15 to 1.47. The following variations of simulated-to-experimental agreements along time range are observed: 1. <sup>nat</sup>Pb and <sup>209</sup>Bi: the simulated activities during the first 15 hours after irradiation are on the average underestimated by a factor of 2, meanwhile, the results are in a much better agreement later. 2. <sup>63</sup>Cu, $`E_p=1186`$ MeV: the simulated results, both for activity and dose rate, are in average 3 times underestimated during the first 5-300 hours after irradiation, meanwhile the discrepancies are negligible for other times. 3. <sup>63</sup>Cu and <sup>65</sup>Cu, $`E_p=127`$ MeV: the simulated results, both of activity and dose rate, are in average 1.5 times lower for <sup>63</sup>Cu, and 2 times underestimated for <sup>65</sup>Cu after $``$ 1 day after irradiation, meanwhile during the first day after irradiation the results either practically coincide with experimental data (<sup>63</sup>Cu) or lie very close to them (<sup>65</sup>Cu). The largest discrepancy between simulated and experimental data is observed for <sup>59</sup>Co irradiated with 127 MeV protons. Simulated activities are about 4 times, and dose rates – 2 times underestimated almost for the entire range of time. This reflects, probably, the importance of nuclear structure effects for this reaction (relatively low bombarding energy and vicinity of the target <sup>59</sup>Co to the doubly magic <sup>56</sup>Ni and magic <sup>57</sup>Ni nuclides) and an inadequate capability of LAHET to describe nuclide production near closed shells, which could be a result of using not good enough approximations for level densities, for inverse cross sections of evaporated and preequilibrium particles, and/or, for nuclear masses and binding energies near closed shells. CONCLUSIONS As a whole, the simulated activities and dose rates are in good agreement with experimental data for most of the targets along the whole time range. Nevertheless, some serious discrepancies at certain times after irradiation, especially for <sup>59</sup>Co, were obtained. At the moment, we have not made a detailed investigation of the causes of the observed discrepancies. The list of products that are produced in the irradiations and/or are observed at decay phase and much contributes to the discrepancies observed have not been defined yet. Also, the contribution of various modes of nuclides formation in nuclear interactions (spallation, fission, fragmentation) to the observed discrepancies has not been analyzed properly. This work is our first step in the study of experimental and simulated activities of targets induced by intermediate and high energy protons. The analysis of observed discrepancies will be continued, and, if possible, will be extended to thick targets in which secondary neutrons, rather than the incident protons, have the largest contribution to target activation. The use of the ISABEL model in LAHET for the higher-energy simulations above its nominal 1-GeV limit could lead to inconsistencies; similar calculations using the alternate Bertini model, or preferably the nuclear models of the CEM95 code , in LAHET are called for. The experimental part of this work have been supported by the ISTC Project #839 and the theoretical study was supported by the U. S. Department of Energy under contract W-7405-Eng-36. REFERENCES 1. Yu. E. Titarenko et al., Nucl. Instr. and Meth. A 414, 73 (1998). 2. Yu. E. Titarenko et al., Proc. Second Int. Topical Meeting on Nuclear Applications of Accelerator Technology (AccApp’98), Gatlinburg, TN, USA, September 20-23, 1998, pp. 164-171 and references therein. 3. R. E. Prael and H. Lichtenstein, “User guide to LCS: The LAHET Code System,” Los Alamos National Laboratory Report LA-UR-89-3014 (1989). 4. W. B. Wilson, T. R. England, and K. A. Van Riper, “Status of CINDER’90 Codes and Data,” Proc. Fourth Workshop on Simulating Accelerator Radiation Environments (SARE4), Knoxville, TN, USA, September 14-16, 1998, p. 69. 5. N. G. Gusev, V. P. Mashkovich, G. V. Obvintsev, “Gamma Irradiation of Radioactive Isotopes and Fission Products,” Moscow (1958) \[in Russian\]. 6. S. G. Mashnik, “User Manual for the Code CEM95”, JINR, Dubna, 1995; http://www.nea.fr/abs/html/iaea1247.html.
no-problem/9908/nucl-th9908055.html
ar5iv
text
# Nucleon-Nucleon Scattering and Effective Field Theory: Including Pions Non-perturbatively ## I Introduction There has been much recent interest in applications of the chiral perturbation theory approach to processes involving an arbitrary number of nucleons introduced by Weinberg . For processes involving more than one nucleon Weinberg’s power counting is applied to the potential rather than to the scattering amplitude. For $`n`$-nucleon processes the potential is defined as a sum of $`n`$-nucleon irreducible time-ordered perturbation theory diagrams. The amplitude is obtained by solving Lippmann-Schwinger equation (or Schrödinger equation). Iteration of the potential via the Lippmann-Schwinger equation leads to divergences. One could try to regularize the potential and include counter-terms, but due to the non-renormalizability of the theory one would have to include an infinite number of (counter-)terms with more and more derivatives. Hence, one has either to exactly solve the equation and after subtract divergences explicitly, or otherwise draw all relevant diagrams, subtract them and sum up. There are not any equations for subtractively renormalised amplitudes. Fortunately there exist practically more powerful approach, cut-off EFT . Divergences appearing in diagrams can be regulated using (sharp or smooth) cut-off regularization. One can keep cut-off parameter of the order of the mass of lightest integrated particle and fit coupling constants to the experimental data as was done in ref. . Cut-off EFT is equivalent to subtractively renormalised EFT up to the order of considered accuracy . Technically more convenient approach based on a new systematic power counting has been suggested by Kaplan, Savage and Wise (KSW counting) in . Pions are included perturbatively within this new scheme. A similar power counting with perturbative pions has been suggested by Lutz . As was mentioned in the expansion parameter in KSW counting is rather large. It raises a question of the usefulness of suggested expansion. Arguments supporting as well as criticising the perturbative inclusion of pions have been given in the literature (see and citations therein. also contains references to lots of papers devoted to EFT study of $`NN`$ interaction problem). In Weinberg’s power counting the one pion exchange potential is of leading order and hence it has to be iterated via the Lippmann-Schwinger equation. KSW counting suggests that contributions containing multiple iterations of the one-pion exchange potential are higher order and hence need not be included in low order calculations. In this letter the results of the application of subtractive renormalization procedure to the Weinberg’s approach to $`{}_{}{}^{1}S_{0}^{}`$ $`NN`$ scattering problem are briefly discussed (the details will be given in separate paper). Leading order (contact interaction plus one-pion exchange) plus next-to-leading order contact interaction potential is iterated using Lippman-Schwinger equation. According Weinberg’s power counting the leading order potential for $`{}_{}{}^{1}S_{0}^{}`$ $`NN`$ scattering is the following: $$V_0(𝐩,𝐩^{})=\stackrel{~}{C}+V_\pi (𝐩,𝐩^{})$$ (1) where $$\stackrel{~}{C}C+\frac{g_A^2}{2f_\pi ^2},V_\pi (𝐩,𝐩^{})\frac{4\pi \alpha _\pi }{\left(𝐪^2+m_\pi ^2\right)},\alpha _\pi \frac{g_A^2m_\pi ^2}{8\pi f_\pi ^2}.$$ (2) $`C`$ is determined by contact interaction terms in the effective Lagrangian , $`𝐪=𝐩𝐩^{}`$, $`g_A=1.25`$ is the axial coupling constant, $`m_\pi =140\mathrm{MeV}`$ is the pion mass and $`f_\pi =132\mathrm{MeV}`$ is the pion decay constant. Substituting $`V_0`$ into the Lippman-Schwinger equation one finds for the Feynman amplitude : $$iA=iA_\pi i\frac{\stackrel{~}{C}\left[\chi _𝐩\right]^2}{1\stackrel{~}{C}G_E}$$ (3) where the quantities $`A_\pi `$, $`\chi _𝐩`$ and $`G_E`$ correspond to diagrams drawn in FIG. 1. $`A_\pi `$ and $`\chi _𝐩`$ are finite. The first diagram contributing to $`G_E`$ is linearly and the second logarithmically divergent, all other diagrams are finite. As $`A`$ contains divergences it is necessary to regularize and renormalize it. Let us introduce a cut-off into the potential: $$V_0^\mathrm{\Lambda }(𝐩,𝐩^{})=\frac{\stackrel{~}{C}\mathrm{\Lambda }^8}{\left(𝐩^2+\mathrm{\Lambda }^2\right)^2\left(𝐩_{}^{}{}_{}{}^{2}+\mathrm{\Lambda }^2\right)^2}+V_\pi (𝐩,𝐩^{})$$ (4) where $`\mathrm{\Lambda }`$ is a cut-off parameter. Substituting $`V_0^\mathrm{\Lambda }`$ into the Lippmann-Schwinger equation one gets the following expression for the amplitude: $$iA^\mathrm{\Lambda }=iA_\pi i\frac{\stackrel{~}{C}\left[\chi _𝐩^\mathrm{\Lambda }\right]^2}{1\stackrel{~}{C}G_E^\mathrm{\Lambda }}$$ (5) where $`A_\pi `$ is unchanged, $`\chi _𝐩^\mathrm{\Lambda }`$ and $`G_E^\mathrm{\Lambda }`$ are given by the same diagrams in FIG.1 with $`\stackrel{~}{C}`$ replaced by $`\frac{\stackrel{~}{C}\mathrm{\Lambda }^8}{\left(𝐩^2+\mathrm{\Lambda }^2\right)^2\left(𝐩_{}^{}{}_{}{}^{2}+\mathrm{\Lambda }^2\right)^2}`$. Note that $`\chi _𝐩^\mathrm{\Lambda }=\chi _𝐩+0(1/\mathrm{\Lambda })`$ and $`G_E^\mathrm{\Lambda }=a_1\mathrm{\Lambda }+a_2\mathrm{ln}\mathrm{\Lambda }+G_E^f+0(1/\mathrm{\Lambda })`$, where $`G_E^f`$ is finite and $`\mathrm{\Lambda }`$-independent. It is straightforward to find $`a_1`$ and $`a_2`$ by calculating analytically first two diagrams contributing to $`G_E^\mathrm{\Lambda }`$. One can numerically calculate $`A_\pi `$ and $`A^\mathrm{\Lambda }`$ for different values of $`\mathrm{\Lambda }`$. Using these values of $`A_\pi `$ and $`A^\mathrm{\Lambda }`$ one finds: $$\frac{1}{A^\mathrm{\Lambda }A_\pi }=B_1(p)\mathrm{\Lambda }+B_2(p)\mathrm{ln}\mathrm{\Lambda }+B_3(p).$$ (6) Although $`B_3(p)`$ actually depends on $`\mathrm{\Lambda }`$, for very large $`\mathrm{\Lambda }`$ it practically does not and $`B_1(p)`$ and $`B_2(p)`$ can be calculated very accurately. Having calculated $`a_1`$, $`a_2`$, $`B_1(p)`$ and $`B_2(p)`$ one easily finds: $$\left[\chi _𝐩^\mathrm{\Lambda }\right]^2=\frac{a_1}{B_1(p)}$$ (7) $$\left[\chi _𝐩^\mathrm{\Lambda }\right]^2=\frac{a_2}{B_2(p)}$$ (8) Good agreement between numerical results obtained from (7) and (8) suggests that numerical calculations are enough accurate. $`\left[\chi _𝐩\right]^2`$ obtained this way is plotted in FIG. 2. The same quantity was calculated in using a different method. The agreement between FIG. 2 and the results of is quite satisfactory. Having calculated $`\left[\chi _𝐩^\mathrm{\Lambda }\right]^2\left(\left[\chi _𝐩\right]^2\right)`$ it is not difficult to calculate the sum $`\stackrel{~}{G}_E^\mathrm{\Lambda }\left(\stackrel{~}{G}_E\right)`$ of all finite regularised (unregularised) diagrams contributing to $`G_E^\mathrm{\Lambda }\left(G_E\right)`$. Phase shifts corresponding to subtractive renormalization procedure can be calculated performing following steps: 1. Calculate $`A_\pi `$, calculate $`A^\mathrm{\Lambda }`$ for very large $`\mathrm{\Lambda }`$ ($`\mathrm{\Lambda }10^5\mathrm{MeV}`$ has been taken); 2. Calculate $$W(p)=\frac{1}{\stackrel{~}{C}}+\stackrel{~}{G}_E^\mathrm{\Lambda }=\frac{\left[\chi _𝐩^\mathrm{\Lambda }\right]^2}{A^\mathrm{\Lambda }A_\pi }I^\mathrm{\Lambda }(p)I_\pi ^\mathrm{\Lambda }(p)$$ (9) where $`I^\mathrm{\Lambda }(p)`$ and $`I_\pi ^\mathrm{\Lambda }(p)`$ correspond to the first and second divergent diagrams contributing to $`G_E^\mathrm{\Lambda }`$ respectively. 3. Calculate two divergent diagrams contributing to $`G_E`$ analytically, subtract the second logarithmically divergent diagram at $`p^2=\mu ^2`$ and the first linearly divergent diagram at $`p^2=\nu ^2(=\kappa \mu ^2)`$ (one is free to subtract these two diagrams at different points). In calculations of this work $`\mu =240\mathrm{MeV}`$ and $`\nu =800\mathrm{MeV}`$ were taken ($`\kappa =100/9`$). Note that the large value of the second subtraction point (800 MeV) does not affect the power counting. Subtracted diagrams give: $$I_\pi ^R(p)=\frac{g_A^2}{2f^2}\left(\frac{m_\pi M}{4\pi }\right)^2\left[i\mathrm{tan}^1\left(\frac{2p}{m_\pi }\right)\frac{1}{2}\mathrm{ln}\left(1+\frac{4p^2}{m_\pi ^2}\right)+\mathrm{ln}\left(1+\frac{2\mu }{m_\pi }\right)\right]$$ (10) and $$I^R(p)=\frac{M(\nu +ip)}{4\pi }$$ (11) 4. Calculate renormalized amplitude: $$A^R=A_\pi +\frac{\left[\chi _𝐩^\mathrm{\Lambda }\right]^2}{W(p)+I^R(p)+I_\pi ^R(p)}$$ (12) 5. Fit $`C`$ so as to reproduce phase shifts for very low energies ($`p5\mathrm{MeV}`$). One gets $`C=1/(17348\mathrm{MeV}^2)`$. 6. Calculate phase shifts. Comparing phase shifts calculated for $`\mathrm{\Lambda }=1\times 10^5\mathrm{MeV},2\times 10^5\mathrm{MeV},3\times 10^5\mathrm{MeV},4\times 10^5\mathrm{MeV}`$ one sees that $`A^R`$ is practically cut-off independent for very large $`\mathrm{\Lambda }`$. These leading order phase shifts are plotted in FIG. 3. As was argued in $`C_2`$ coupling constant of the part of the next-to-leading order potential $$V_1=C_2\left(𝐩^2+𝐩_{}^{}{}_{}{}^{2}\right)/2$$ (13) is larger than assumed by Weinberg’s power counting. This suggests that contributions of this part of the next-to-leading order potential are larger than of others. This observation simplifies the actual next-to-leading order calculations significantly. It is possible to sum up all diagrams obtained by iterating Lippman-Schwinger equation for the potential $`V=V_0+V_1`$, where $`V_0`$ and $`V_1`$ are given by (1) and (13) respectively. The result obtained in using dimensional regularization reads: $$A_1(𝐩)=A_\pi \frac{\left[\chi _𝐩\right]^2}{\left[\stackrel{~}{C}\alpha _\pi m_\pi MC_2+C_2𝐩^2\right]^1G_E}$$ (14) As $`\chi _𝐩`$ and $`\stackrel{~}{G}_E`$ are already calculated (although these quantities were calculated using cut-off regularization, as was mentioned above the effect of finite cut-off is extremally small and hence negligible), it is straightforward to apply the same subtraction scheme to (14), fix $`C_2`$ so as to fit the data for $`p=10\mathrm{MeV}`$ and calculate the phase shifts. The results are plotted in FIG. 3. One can see that agreement with Nijmegen phase shift data is quite good. Obtained value $`C_2=1/(8.5\times 10^9\mathrm{MeV}^4)`$ is indeed larger than assumed by Weinberg’s counting. One also can include $`V_1`$ perturbatively expanding (14) in $`C_2`$ and retaining only first two terms. Calculated phase shifts are also plotted in FIG. 3. Comparing the phase shifts for perturbative and non-perturbative inclusion of $`V_1`$ one sees that deviation is small for momenta up to 300 MeV. This suggests that higher order contributions of $`V_1`$ are small as is expected from power counting arguments. On the other hand it is straightforward to include pions perturbatively (as was done in ) following ideas of KSW counting. Using Those values for $`C`$ and $`C_2`$ which were obtained above one gets negative phase shifts. this suggests that the higher order contributions of the one pion exchange potential are by no means small. Note that subtraction scheme used here is different from the one considered in . In finite diagrams were subtracted as well. Repeating the calculations of this work with non-perturbatively included one pion exchange potential using PDS scheme and substituting the “best fit” parameters from one gets again negative phase shifts. This comparison of the phase shifts calculated using perturbative and non-perturbative inclusion of pions supports the conclusion of that the consistency of the perturbative inclusion of pions in EFT approach to the $`NN`$-scattering problem is questionable. ACKNOWLEDGEMENTS This work was carried out whilst the author was a recipient of an Overseas Postgraduate Research Scholarship and a Flinders University Research Scholarship at the Flinders University of South Australia.
no-problem/9908/hep-lat9908054.html
ar5iv
text
# On the phase diagram of 2d Lorentzian Quantum Gravity ## 1 INTRODUCTION Recently a new model of 2d quantum gravity has been proposed . It is defined using dynamical triangulations from a subclass of diagrams which can be given a causal structure. Such diagrams are generated by gluing together one dimensional time–slices or “universes” (in our case a set of vertices connected by space-like links forming a diagram with the topology of a circle) with time-like links such that they form a triangulated surface. Vertices connected by time-like links are causally related and a unique time can be assigned to the vertices of each time–slice. Such graphs can be given a Lorentzian metric by defining time-like links to have equal negative length squared and space-like to have positive. All triangles have equal area and the volume of spacetime is proportional to the number of triangles $`N_T`$. The system has been found to have a non–trivial continuum limit only at an imaginary value of the cosmological constant $`\lambda `$. The geometry of space is maximally fluctuating but the system is much smoother than Liouville gravity: By defining the two point function to be $$G(\lambda ,t)=\underset{T𝒯}{}\text{e}^{\lambda N_T},\lambda \mathrm{I}\mathrm{R},$$ (1) where the summation is over triangulations $`T`$ of cylindrical topology with $`t`$ time–slices, one finds that the Hausdorff dimension $`d_H`$ of the system is $`2`$. This is entirely due to the imposition of the causal structure: If one allows the creation of baby universes the system becomes the ordinary EQG model. If conformal matter of central charge $`c`$ is coupled to the system, it has been found in the case of one Ising model that the coupling is weak for $`c1/2`$ and the bulk properties of geometry do not change (e.g. $`d_H=2`$). The critical matter system belongs to the Onsager universality class. When 8 Ising spins are coupled to gravity, a qualitatively different behaviour emerges. The system undergoes a phase transition for some $`1/2<c_{\mathrm{crit}}<4`$ to a phase where we observe anomalous scaling between the typical length scale $`L`$ and typical time scale $`T`$. It is found that $`\mathrm{dim}[L]=2\mathrm{dim}[T]`$ and that $`d_H=3`$. The latter is a “cosmological” Hausdorff dimension. The former is a relation which has been found to hold for the non–singular part of the quantum geometry of EQG for any value of $`c1`$. The short distance behaviour of space–time is given by a different fractal dimension $`d_h=2`$. The quantum geometry is different at different scales and has a more complex structure. The matter critical exponents are Onsager. We conclude that the matter coupling to geometry in our model is much weaker than in the case of EQG so that the large $`c`$ phase has non–degenerate, interesting continuum limit for the quantum geometry. Nothing dramatic happens from the point of view of matter but the geometry undergoes a qualitative change and even shares some features with the non–singular part of EQG in the $`c<1`$ phase. This is the first time where one has a model which demonstrates explicitly that the strong coupling of matter and geometry in Liouville gravity is entirely due to the (a priori) presence of baby–universes. ## 2 NUMERICAL RESULTS We simulate triangulated surfaces with the topology of the torus with $`N_T`$ triangles and $`N=N_T/2`$ vertices . The temporal length (number of time slices) is $`t`$ and we choose $`N=t^2/\tau `$ for $`\tau =1,2,3`$ and $`4`$. The geometry is updated with the move described in and the Ising spins with the Swendsen–Wang algorithm. The partition function for $`n`$ Ising models is $$G(\lambda ,t,\beta )=\underset{T𝒯}{}\mathrm{e}^{\lambda N_T}Z_T^n(\beta ).$$ (2) $`Z_T(\beta )`$ is the partition function of an Ising model with spins placed at the vertices of the graph $`T`$ at inverse temperature $`\beta `$. $`n=1,8`$ corresponding to $`c=1/2,4`$. We obtain the fixed volume partition function by adding a gaussian volume–fixing term and measuring only on configurations of given volume $`N_T`$. The critical values of $`(\lambda ,\beta )`$ are determined. For $`c=0`$ they are $`\lambda _c=\mathrm{ln}2`$, for $`c=1/2`$ they are $`(\lambda _c,\beta _c)=(0.742(5),0.2521(1))`$ and for $`c=4`$ they are $`(\lambda _c,\beta _c)=(1.081(5),0.2480(4))`$ ($`\lambda \mathrm{ln}2`$ when $`\beta \mathrm{}`$). Those values are insensitive to $`\tau `$. Finite size scaling (FSS) is applied at $`(\lambda _c,\beta _c)`$ in order to measure the system’s scaling properties. The first quantity that we measure is the distribution $`SV(l)`$ of spatial volumes $`l`$. We expect a scaling behaviour $$SV_N(l)=F_S(l/N_T^{1/\delta _h}),$$ (3) for some function $`F_S`$. $`\delta _h`$ is related to the fractal dimension $`tN_T^{1/d_H}`$ by $`d_H=\delta _h/(\delta _h1)`$. Here, $`t`$ is the dynamically generated time extend of the scaling part of space–time. For the $`c=1/2`$ model we find $`\delta _h=d_H=2`$ as shown in Fig. 1. For $`c=4`$ the scaling behaviour of $`SV(l)`$ is qualitatively different. As can be seen in Fig. 3, the configurations show a tendency to form a long and thin neck with spatial volume of the order of the cutoff and an extended region that scales according to (2). The volume of the extended part is asymptotically proportional to $`N_T`$. The effect is seen for large enough volumes and finite size effects are minimised by taking $`\tau 3`$. For $`\tau =3`$ we find $`\delta _h=1.54(3)`$ (see Fig. 2) and for $`\tau =4`$ $`\delta _h=1.50(3)`$. Let us assume for clarity that $`\delta _h=3/2`$. Then $`d_H=3`$ and $`lt^2`$ where $`l`$ is the typical length scale of universes in the extended region and $`t`$ the typical time scale of their existence. The above results are supported by measuring the volumes $`n_N(r)`$ of geodesic spherical shells at distance $`r`$. We expect that $`n_N(r)=N_T^{11/d_H}F_1(x)`$, $`x=r/N_T^{1/d_H}`$ which is known to hold in EQG for all $`c`$. $`d_H`$ is the cosmological Hausdorff dimension describing how volume and time have to scale in order to obtain a non trivial continuum limit of the two loop function. On the other hand the short distance behaviour is given by $`d_h`$, where $`n_N(r)r^{d_h1}`$ at scales $`rN^{1/d_H}`$. For $`c=0`$ we find analytically that $`d_H=d_h=2`$ and for $`c=1/2`$ numerically that $`d_H=d_h=2.00(5)`$. For $`c=4`$, $`n_N(r)`$ has different scaling behaviour at different length scales. For $`rN^{1/3}`$ we find that $`d_h=2.1(2)`$ as can be seen of Fig. 4. For $`rN^{1/3}`$ the value of the tail of $`n_N(r)`$ is almost independent of $`N_T`$ and $`r`$ showing dominance of 1d configurations. For $`rN^{1/3}`$ the scaling of the peaks of $`n_N(r)`$ gives $`d_H=3.07(9)`$. The matter scaling exponents are computed from the scaling behaviour of the magnetic susceptibility $`\chi N^{\gamma /\nu d_H}`$ and $`D_{\mathrm{ln}|m|}\frac{d\mathrm{ln}|m|}{d\beta }N^{1/\nu d_H}`$ and $`D_{\mathrm{ln}m^2}\frac{d\mathrm{ln}m^2}{d\beta }N^{1/\nu d_H}`$. The $`c=1/2`$ system clearly belongs to the Ising universality class. For $`c=4`$ special care has to be taken in order to isolate the critical behaviour of the spin system in the extended region of space–time . The critical exponents are consistent with Onsager values. Our results are summarized in Table 1.
no-problem/9908/nucl-th9908037.html
ar5iv
text
# Generalized seniority from random Hamiltonians ## I Introduction Pairing in fermion systems is a ubiquitous phenomenon that appears among fermions as diverse as electrons, nucleons, and <sup>3</sup>He atoms. In the nuclear shell model this is usually explained as due to strong, attractive matrix elements of the two-body effective interaction between $`J=0`$ pair states. In the Fermi liquid model it is explained as a consequence of the interaction being attractive at certain momentum and energy transfers. In this paper we wish to demonstrate that features of pairing arise from a very large ensemble of two-body interactions and, hence, are independent, to a large extent, from the specific character of the interaction. Pairing may be favored simply as a consequence of the two-body nature of the interaction and the way it connects the Fock space wave functions of the noninteracting Fermion system. We examine this possibility by numerical studies of the many-body system governed by a two-body Hamiltonian taken from a random ensemble. We consider nucleons in a spherical shell-model space for which $`J=0`$ states have a special significance. We thus take ensembles of two-body Hamiltonians that respect angular momentum, but otherwise are as general as possible. In an earlier paper, we examined spectral features of pairing with one such ensemble and found that two signatures were present in the preponderance of Hamiltonians: the ground state tends to have $`J=0`$ angular momentum, and there tends to be a gap between that state and the higher states in the spectrum. Another important feature shown in was referred to as phonon collectivity. Ground states with $`J=0`$ had, on the average, rather large matrix elements of a single-nucleon operator with the first excited $`J=2`$ states. This is a spectral characteristic of pairing, but it may occur in other situations as well. Why pairing should be favored in a general ensemble is not clear. It has been suggested that time-reversal invariance is responsible, but that symmetry has been found to be unnecessary . In any case, energy spectra are only part of the properties affected by pairing. The term “pairing” implies that the ground state can be approximated or modeled by a condensate of pairs of fermions coupled, in the interacting shell model, to zero angular momentum; that is, the ground state is approximately of the form $`(_{\alpha ,m}c_\alpha a_{\alpha ,m}^{}a_{\alpha ,m}^{})^N|0`$. Furthermore, under some conditions the matrix elements of pairing operators have an algebraic structure, as described by the seniority model and its generalizations (see ). Along these lines, we consider here two signatures that explicitly probe the wave functions: $``$ a strong pair-transfer amplitude; that is, there is a large matrix element of the pair annihilation operator between the ground states of the nuclei with $`A2`$ and $`A`$ nucleons; $``$ considering an isotopic chain of nuclides together, $`A`$, $`A2`$, $`A4`$, etc., the seniority model and its generalizations predict that the same pair annihilation operator that takes one from the $`A`$ ground state to the $`A2`$ ground state will also take one from the $`A2`$ to the $`A4`$ ground state, etc. We also consider here an additional energy signature: $``$ in an isotopic chain containing both even and odd $`A`$, the even isotopes have systematically greater binding energies. Besides considering these three additional signatures of pairing, we will examine some other ensembles to see how robust our results are. The calculations are performed in two shell model spaces: the “$`sd`$” space consisting of the orbitals with angular momentum $`j=1/2,3/2,5/2`$ which can accommodate up to 12 identical particles, and the “$`pf`$” space which has in addition the $`j=7/2`$ shell and accommodates up to 20 identical particles. We use the Glasgow-Los Alamos and ANTOINE shell-model codes to calculate the many-body wave functions and observables in these shell-model spaces. The Hamiltonian is specified by the single-particle and two-particle matrix elements. Except for one ensemble discussed below, we set the single-particle matrix elements to zero. ## II Two-Body Random Interactions For the two-body matrix elements, we choose a basis of two-body states, labeled by $`\alpha `$, which has good angular momentum $`J`$. There are 63 independent two-body matrix elements in the $`sd`$ space and 195 in the $`pf`$ space, including both neutrons and protons. We define an ensemble of two-particle Hamiltonians requiring that the ensemble be invariant under changes in the basis of two-particle states. This can be achieved by taking the matrix elements to be Gaussian distributed about zero with the widths possibly depending on $`J`$, $$V_{\alpha ,\alpha ^{}}^2=c_{J_\alpha }(1+\delta _{\alpha ,\alpha ^{}})\overline{v}^2.$$ (1) $$V_{\alpha ,\alpha ^{}}V_{\beta ,\beta ^{}}=0,(\alpha ,\alpha ^{})(\beta ,\beta ^{})$$ Here $`\overline{v}`$ is an overall energy scale that we generally ignore (except for scaling single-particle energies for the RQE-SPE defined below). The coefficients $`c_J`$ then define the ensemble. We emphasize that $`J`$ refer to quantum numbers of two-body states, and not of the final many-body states (typically 4-10 particles). We now discuss the choices of ensembles, which may be specified by the $`c_J`$ coefficients and the single-particle Hamiltonian, if present. In our earlier work we employed the ‘RQE’ ensemble defined below, but it is important to examine other ensembles to see how robust the results are. The ensembles are: 1. RQE (Random Quasiparticle Interaction). Here $`c_J=(2J+1)^1`$. The relation between the $`c_J`$ came from imposing an additional invariance on the ensemble, that it be the same for the particle-particle interaction as for the particle-hole interaction. The RQE gives a larger variance to $`J=0`$ matrix elements than to the others. Note that even though the variance is larger, the matrix elements are both attractive and repulsive, so there is no bias toward pairing by the traditional mechanism of an attractive two-particle interaction. 2. TBRE (Two-body Random Ensemble). Here $`c_J=`$ constant. Historically, this was the first two-particle random ensemble to be employed in studying statistical properties of many-particle spectra . 3. RQE-NP (Random Quasiparicle Ensemble-No Pairing). This is the same as the RQE ensemble, except all $`J=0`$ two-body matrix elements are set equal to zero. This ensemble will show clearly whether the $`J=0`$ channel matrix elements are needed at all to produce the signatures of pairing. (It is known , albeit not widely appreciated, that it is possible to have interactions that are diagonal in seniority without any explicit $`J=0`$ pairing interaction.) 4. RQE-SPE (Random Quasiparticle Ensemble-with Single-Particle Energies). All the previous ensembles had the single-particle energies set to zero. Realistic interactions do have nonzero single-particle energies, and these can, in principle, affect pairing properties, at the very least by creating large shell gaps. For calculations in the sd shell we take here single-particle energies from the Wildenthal interaction , scaling $`\overline{v}=3.84`$ MeV so as to best match the widths of the two-particle matrix elements. For the pf shell we use single particle energies from the modified KB3 interaction and scaled $`\overline{v}=4.43`$ MeV. ## III Results For the specific calculations, we considered 4, 6, and 8 neutrons in the $`sd`$ space; we label these as the corresponding shell-model systems, <sup>20</sup>O, <sup>22</sup>O, and <sup>24</sup>O respectively; we caution the reader that this labeling can be misleading as we have deliberately put in as little physics of those systems as possible. We also considered 4, 6, 8, and 10 neutrons in the $`pf`$ space: <sup>44</sup>Ca, <sup>46</sup>Ca, <sup>48</sup>Ca, and <sup>50</sup>Ca. Finally, we included systems with nontrivial isospin, considering in the $`sd`$-shell 4 protons and 4, 6, and 8 neutrons: <sup>24,26,28</sup>Mg, respectively. In these cases $`c_Jc_{JT}`$. Thus for the RQE, $`c_{JT}=(2J+1)^1(2T+1)^1`$. For each of these systems, and for each of the ensembles described above, we computed at least 1000 samples. Spectral signatures. Table 1 presents the fraction of each ensemble that yields a $`J=0`$ ground state for the above systems. For purposes of comparison, the fraction of the total many-body states that are $`J=0`$ and $`J=2`$ states is also given. If the ground state spins reflected only the size of the $`J`$ subspace, there would be more $`J=2`$ than $`J=0`$, contrary to our findings. In addition to a predominance of $`J=0`$ ground states, such states are pushed down relative to the rest of the spectrum. An example is shown in Fig. 1. Note that for spectrum 1(a), the $`J=0`$ ground state is separated from the excited states by an amount large compared to the average level spacing, while for the case of a $`J>0`$ ground state, 1(b), the separation of the ground state and the average level spacing is similar. This is shown in more detail in Fig. 2. Here we define $`s`$ to be the spacing between the ground state and the first excited state, scaled by the local level spacing $`D`$, defined as the ensemble-averaged spacing between the first and second excited states. Because these states in general do not have the same quantum numbers such as total $`J`$, one would expect the level spacing to be described by a Poisson distribution, where the probability of finding a spacing $`s`$ is given by $`P(s)=\mathrm{exp}(s/D)`$ . For cases where the ground state $`J0`$, the Poisson distribution describes the distribution of $`s`$ extremely well, as one would predict. For those cases where the ground state $`J=0`$, however, the distribution is much broader. It is somewhat approximated by a Poisson, but with $`D`$ 3 times larger. We show two cases in Fig. 2. The other nuclides and ensembles yield nearly identical figures. Table II tabulates the average $`s`$ for the various ensembles. For all cases, a $`J=0`$ ground state is pushed down an average factor of 2.3-3.7 relative to the local level spacing, whereas a $`J>0`$ ground state is, within statistics, not pushed down at all. Similar results hold for the Mg ensembles. The third spectral feature is the well-known even-odd staggering of ground state energies. Figures 3 illustrates the real world situation with the experimental neutron removal energies $`S_n(A)=E(A)+E(A1)`$ of calcium isotopes in the range A=45-50. The larger removal energy of the even isotopes is associated with their greater binding energy. We look for evidence of this in our ensemble spectra of the $`pf`$ isotope chains $`A=410`$ as follows. We first examined the even members of the chain, requiring that all ground states have $`J=0`$. This is satisfied for $`42\%`$ of the members of the RQE ensemble; this is a much larger than expected value $`(0.70)^4=0.25`$ that one would obtain from Table 1 assuming that the $`J=0`$ occurrences are uncorrelated. In the generalized seniority model , the even-member ground state energies have a quadratic dependence on $`A`$, eq. (23.20): $$E_{gs}(A)=a+bA+cA^2.$$ (2) We next make a least-squares fit of the selected even-$`A`$ chains to this formula. This is, of course, a fit of 3 parameters to 4 data points and the description is good. Examples of the deviations about this fit are plotted in Fig. 4. We then computed the binding energies for $`n=5,7`$. The deviations from (2) for <sup>47,48</sup>Ca are plotted in Fig. 4, scaled to the local level spacing: <sup>48</sup>Ca exemplifies all the even-$`n`$ cases, which are all very similar, while <sup>45</sup>Ca yields a plot nearly identical to that of <sup>47</sup>Ca. Notice that not only are the odd-particle systems consistently higher in energy, they are pushed up on average by 3 times the local level spacing – which is entirely consistent with the results shown in Fig. 2 and Table II. Figure 4 also contains results from the RQE-NP ensembles in the $`pf`$ shell. Even with all $`J=0`$ matrix elements set to zero, we find qualitatively similar results. The effects are not as dramatic in this case; from Table I one would expect all four isotopes <sup>44,46,48,50</sup>Ca to have $`J=0`$ ground states $`6.5\%`$ of the time, but in fact this occurs $`8.4\pm 0.8\%`$ of the time. Pair-transfer collectivity. The spectral and energetic characteristics discussed above are not the only signatures of pairing; matrix elements of pairing operators are also very important. In order to test the hypothesis that the ground states of these random Hamiltonians can be approximated by pair condensates, let’s follow the example of generalized seniority and consider the general pair-annihilation operator $`S=_j\alpha _jS_j`$, where $`S_j=_{m>0}()^ma_{jm}a_{jm}`$ is the pair-annihilation operator for the $`j`$-shell. Given $`S`$, the pair-transfer amplitude from the ground state with $`A`$ particles to $`A2`$ particles is $`A2|S|A`$. One way to probe the wave function is the pair-transfer fractional collectivity (defined in analogy with the phonon fractional collectivity of Ref. ): $$f_p=\frac{A2|S|A^2}{A|S^{}S|A}.$$ (3) If the states of the system are condensates of the $`S^{}`$ pairs, then one expects $`f_p=1`$. How does one determine the $`\alpha _j`$? Because the ensembles are defined to be invariant on changes of basis, there cannot be a globally preferred $`\alpha _j`$. In principle, we could determine individual $`\alpha _j`$ for each ensemble member by maximizing $`f_p`$ from Eq. (3). However, the variational condition is rather complicated, and we found satisfactory evidence of pairing collectivity with a much simpler ansatz. In analogy to phonon fractional collectivity used in , we set $$\alpha _j=A2|S_j|A.$$ (4) Figure 5 presents the distribution of the $`f_p`$ for various ‘nuclides’ and interaction ensembles. The ensemble denoted ‘GOE’ refers to using two different RQE interactions for the $`A`$ and $`A2`$ wave functions; one would expect a minimal correlation between their wave functions and indeed the distribution of $`f_p`$ is heavily weighted towards zero for all nuclides. For the cases using the same interaction for the $`A`$ and $`A2`$ wave functions, however, we get distinctly different results: a weighting towards $`f_p=1`$, implying an enhanced correlation indicative of a pairing-like condensate. All our nuclides and ensembles yield similar plots. The results are summarized in Table III in the form of the average fractional pairing. Keep in mind that the distributions for GOE have a negative slope, while for all other ensembles the slope of the distribution is positive. (We also tabulate, for comparison, the exact $`f_p`$ for realistic interactions: the Wildenthal interaction in the $`sd`$ shell and modified KB3 in the $`pf`$ shell.) Thus, in the cases of Ca and O, for all these ensembles–even those with the $`J=0`$ pairing matrix elements explicitly removed–we see an increased enhanced number of states with a condensate-like ground state. The Mg nuclei lie in between the GOE and ensembles of identical nucleons. This indicates that the proton-neutron interaction dampens the pairing collectivity present in all-neutron systems such as the <sup>20-24</sup>Ca and <sup>44-50</sup>Ca isotopes. The difference is likely due to the $`T=0`$ interaction. For interactions that are truly diagonal in generalized seniority, one expects the same condensate to prevail for $`A=2,4,6,8,\mathrm{}`$ valence nucleons . In the language developed above, let $`\alpha _j(A)`$ be the coefficients computed from $`A`$ and $`A2`$. The $`\{\alpha _j(A)\}`$ can be thought of as vectors, and from generalized seniority we expect the vectors $`\stackrel{}{\alpha }(A)`$ and $`\stackrel{}{\alpha }(A+2)`$ to be aligned. To test this idea, define the scalar product $`\stackrel{}{\alpha }(A)\stackrel{}{\alpha }(A^{})=_j\alpha _j(A)\alpha _j(A^{})`$, where the states and matrix elements are calculated with the same two-body Hamiltonian. (One could have different weightings or metrics for this scalar product, such as $`\sqrt{2j+1}`$ or $`1/\sqrt{2j+1}`$, but such differences in definition do not change our results.) Then plot the distribution of $$\mathrm{cos}\theta =\frac{|\stackrel{}{\alpha }(A)\stackrel{}{\alpha }(A+2)|}{\left|\alpha (A)\right|\left|\alpha (A+2)\right|}$$ (5) Calculations for the O and Ca isotopes using realistic interactions typically give 0.99 for this correlation factor, except at shell closures of the $`d_{5/2}`$ in the $`sd`$ shell and $`f_{7/2}`$ in the $`pf`$ shell, where it is 0.4-0.5. For the Mg isotopes with the Wildenthal interaction, this factor is $`0.7`$, indicating that the likelihood for the same correlated pair to be transferred along the chain is somewhat less than the all-neutron case. If all $`T=0`$ matrix elements are set to zero, then one recovers the factor 0.99 for the correlation. The results for the GOE and RQE ensembles for O and Ca are plotted in Fig. 6. For the GOE case we find a flat distribution – the pair-transfer amplitudes are uncorrelated, exactly as one would expect. However, for the ensembles of random two-body interactions, we find for the O and Ca chains a sharp peak at 1, indicating a strong correlation. The chain <sup>28</sup>Mg$``$<sup>26</sup>Mg$``$<sup>24</sup>Mg, plotted in Fig. 7b, also shows a peak at 1, which is a factor of 4 higher than the average bin height. In contrast, the O and Ca peaks are at least a factor of 10 above the average bin height. Thus, the pair transferred in the O and Ca chains is much more likely to be of the same condensates than the pair transferred in the Mg case. The other ensembles yield plots similar to that shown for the RQE. Curiously enough, for the RQE-NP ensemble (not shown), we also find a sharp peak at $`\mathrm{cos}\theta =0`$, as well as at $`\mathrm{cos}\theta =1`$. The analysis described in the previous paragraph only considered ‘nearest-neighbor’ transitions. If, however, we have an approximate generalized seniority, then we expect the pair-transfer amplitude vector to be similar for a whole chain of isotopes. We compare, for the RQE, the correlation for the pair-transfer amplitudes starting from <sup>50</sup>Ca $`^{48}`$Ca and computing the correlation, not only with <sup>48</sup>Ca $`^{46}`$Ca, but also with <sup>46</sup>Ca $`^{44}`$Ca. This correlation shows an enhancement at the value that is similar to the results shown in Fig. 6. Thus we have strong evidence that the pairing condensate is not an arbitrary and local feature, but persists along an isobaric chain. ## IV Conclusions We have considered several random ensembles of two-body Hamiltonians in the framework of the shell model. By examining the statistical properties of the low-lying spectra, as well as pair-transfer amplitudes, we find pairing behavior occurs frequently in our ensembles of two-body interactions. Thus, pairing is a robust feature of two-body Hamiltonians. There seems to be a large class of two-body interactions leading to pairing which is much wider than the attractive interactions usually considered. Besides pairing, there are other features of nuclear spectra that often occur and can give rise to algebraic structures. The most prominent example, rotational bands, is not favored at all by random Hamiltonians. Since a rotational band implies an internal rigidity of the system, this shows that in some sense the random ensembles describe only Fermi liquid behavior. In the spherical shell model, it has been shown that it is the $`T=0`$ part of the nuclear effective interaction acting between neutrons and protons which gives rise to collective spectra like rotational ones. The effective $`T=0`$ interaction has a rather strong quadrupole component which breaks the seniority coupling scheme. It is interesting to speculate on the more complex algebraic structures that have been found in nuclear spectroscopy. The phenomenologically successful interacting boson model is based on collective pair transfer operators in both $`J=0`$ and $`J=2`$ (quadrupole) angular momenta. Since we see no indication of quadrupole collectivity in the random Hamiltonians, one would have to introduce from the start some physical features of the interaction. It might be that the important physical features could be described very simply, say, by an attractive surface delta interaction. One would then look for the rich variety of observed dynamical symmetries by adding to the physical component a component from one of the random ensembles. ###### Acknowledgements. CWJ and IT thank the Institute for Nuclear Theory and the program on Algebraic Methods in Many-Body Physics, where much of this work was initiated and the first drafts of this paper were written. We had many stimulating discussions, including A. Dieperink, J. Ginocchio, F. Iachello, and S. Tomsovic. This work is supported by Department of Energy grant numbers DE-FG02-96ER40985, DE-FG-06-90ER40561, and DE-FG02-96ER40963. Oak Ridge National Laboratory is managed by Lockheed Martin Energy Research Corp. for the U.S. Department of Energy under contract number DE-AC05-96OR22464.
no-problem/9908/cond-mat9908185.html
ar5iv
text
# Correlated hip motions during quiet standing ## 1 Introduction Physical investigations into the motor control of body motions have traditionally been of three types (Shumway-Cook and Wallacot 1995): (i) direct muscle tension measurements, (ii) kinetic analysis of forces, and (iii) kinematic analysis of body displacements. The experimental techniques used to study body motions reflect the requirements of the above types of investigation. For example, electromyography has been the technique most often used to monitor the activity of muscles (Granley 1984, Winter 1990, Perry et al1981). Electrodes are placed on the skin above the particular muscles of interest. Force transducers and force plates measure the ground body reaction forces to body movements (Prietro et al1993, Firsov et al1993, Winter 1990, Goldie et al1989). In the present work, purely kinematic studies of body motions will be of interest. Previous studies have employed electrical potentiometers to measure joint angles in those cases for which angular increments induce voltage differences (Barlett et al1986, Campbell et al1989). Accelerometers have been constructed from force transducers which also (ultimately) depend on kinematic induced voltages (Thomas and Whitney 1959). Finally, video, cinematography and optoelectric systems have been used to form images of body motions. The optoelectric systems require that infrared sources and/or reflectors be worn on each anatomic landmark to be imaged (Whittle 1991, Winter 1990). Kinematic experimental methods have probed a well known set of strategies that are used by humans to maintain balance (Nasher 1985). The strategies involve the ankles, the hips and stepping. These three forms of motion are exhibited in various degrees depending on the nature of the motor task. The purpose of this work is to introduce a new ultrasonic sound wave assessment (SWA) device which aids the analysis of body movement kinematics. The SWA device measures the coordinates $`X_1,X_2,\mathrm{}`$ of anatomic landmarks as a function of time. This device has been employed to investigate balance strategies during quiet standing. Surprising but unambiguous experimental results show that strongly correlated hip motions play a central role even in the postural sway of an unperturbed stance. This experimental result was made possible due to the accuracy of the SWA device. In Sec.2, the SWA device will be described in detail. In Sec.3, two coordinate measurements are discussed. The measurements involve the random noise motions of quietly standing subjects. Two coordinates were selected so as to emphasize the hip strategy. The observed restricted motions in the two coordinate phase space indicated the dominance of the hip strategy in the quiet postural stance. In Sec.4 we calculate the correlation functions for the two coordinate phase space and deduce that the hip motions are quite strongly correlated in all of the subjects measured. In the concluding Sec.5, previous notions of quiet standing balance strategies will be discussed in the light of presently reported data. ## 2 Sound wave assessment device The SWA device consists of an even number of small ultrasonic transducers. The measurement of $`𝒩`$ coordinates requires $`2𝒩`$ transducers. Each of the $`𝒩`$ pairs of transducers send ultrasonic pulses to one another at a rate of $`1800`$ pulses per minute. One transducer of each pair is positioned on a stable laboratory stand. The other transducer of the pair is attached to a quietly standing subject. Each pulse sent by one transducer in a given pair is later detected by the other transducer in the same given pair. The distance between the two transducers of a given pair can then be obtained from the time required for the pulses to travel from the sender to the receiver. Thus, each pair of transducers measures one coordinate function $`X(t)`$, i.e. the distance between the pair of transducers at time $`t`$. In the photography industry, a single ultrasonic transducer system has been used to measure the distance between the camera lens and the object to be photographed. Twice the distance is computed from the measured pulse emission time and the measured pulse detection time at which the transducer detects its own echo. Our use of transducer pairs improves the accuracy of the coordinate measurement system by having less air absorption as well as less spatial dispersion (fanning out) of the ultrasound signals. Each “pulse” (produced at the rate of $`1800`$ per minute) in reality consists of $`16`$ very closely spaced pulses separated by time intervals of $`\delta t2\times 10^5\mathrm{sec}`$. If a transducer pair is well aligned, then first pulse of the sixteen sub-pulses will be detected. If the transducer pair alignment is somewhat skewed, then the second sub-pulse will be detected. The coordinate displacements are measured to within $`0.02cm`$ in a pulse bandwidth of $`12kHz`$. This yields a displacement noise error of $`\delta X2(\mu m/\sqrt{Hz})`$. Thus, we can record the fine displacements of a subject’s body coordinates $`X_1(t),X_2(t),\mathrm{}`$ as a function of time. Twelve healthy subjects in the age range between $`15`$ and $`65`$ years of age participated in this study. ## 3 Measured paths in two coordinate planes Shown in Fig.1 is the random motion in the $`(X_A,X_M)`$ plane of a subject’s upper body motion during quiet standing. The coordinate $`X_A`$ measures the subject’s upper body anteroposterior displacement, i.e. the forward and backward motions of the transducer attached to the subject’s back. The coordinate $`X_M`$ measures the upper body mediolateral displacement, i.e. the side to side motions of the transducer attached to the subject’s shoulder. The random motions were measured over a time period of one minute. The quantities $`\mathrm{\Delta }X_A=X_A\overline{X}_A`$ and $`\mathrm{\Delta }X_M=X_M\overline{X}_M`$ denote the deviations from the mean taken over the observation time. One notes the meandering nature of the noise in the $`(X_A,X_M)`$ plane. By contrast, in Fig.2, we consider the one minute measurement (on the same subject) in the $`(X_U,X_L)`$ plane with both displacements in the anteroposterior direction. The coordinate $`X_U=X_A`$ was again measured from a transducer attached to the subject’s back (upper body) while the coordinate $`X_L`$ was measured from a transducer attached to the back of the subjects thigh slightly below the hip (lower body). The deviations from the mean are $`\mathrm{\Delta }X_U=X_U\overline{X}_U`$ and $`\mathrm{\Delta }X_L=X_L\overline{X}_L`$. One notes the very constrained nature of the noise in the $`(X_U,X_L)`$ plane. The coordinates $`X_U`$ are highly correlated. The positions in the plane are nearly collinear, as shown in Fig.2 along with a “best fit” line. In Fig.3, we plot both coordinates $`X_U(t)`$ and $`X_L(t)`$ as a function of time for an observation time of one minute. If one views only one of the coordinates, then the motion appears to be random. However when one views both coordinates, the changes in $`X_U(t)`$ and $`X_L(t)`$ are seen to be strongly anti-correlated. When one of the coordinates increases, the other coordinate decreases in proportionate amounts. Thus, when the upper body moves forward, the lower body moves backward and vice versa. This describes precisely the hip balance strategy. ## 4 Statistical analysis A quantitative statistical formulation of the anti-correlated kinematics of the hip strategy may be formulated as follows (Martin 1967): (i) For a given subject, the time averaged coordinate correlation matrix elements may be defined as $`𝒢_{UU}={\displaystyle \frac{1}{\tau }}{\displaystyle _0^\tau }\mathrm{\Delta }X_U(t)\mathrm{\Delta }X_U(t)𝑑t,`$ (1) $`𝒢_{LL}={\displaystyle \frac{1}{\tau }}{\displaystyle _0^\tau }\mathrm{\Delta }X_L(t)\mathrm{\Delta }X_L(t)𝑑t,`$ (2) $`𝒢_{UL}=𝒢_{LU}={\displaystyle \frac{1}{\tau }}{\displaystyle _0^\tau }\mathrm{\Delta }X_U(t)\mathrm{\Delta }X_L(t)𝑑t,`$ (3) where $`\tau =60sec`$ is the observation time. (ii) The normalized cross correlation is defined $$𝒞_{UL}=\frac{𝒢_{UL}}{\left(𝒢_{LL}𝒢_{UU}\right)^{1/2}},$$ (4) which obeys the inequalities $`1𝒞_{UL}+1`$. The extreme of the inequality $`+1`$ indicates perfect correlation, $`1`$ indicates perfect anti-correlation and $`0`$ indicates lack of any correlation. All of the measured subjects exhibited a cross correlation function in the range $`0.89(𝒞_{UL})0.99`$ indicating an almost perfect anti-correlation. For the totality of the experimental data, one may define the ensemble average of the correlation matrix elements for $`N=12`$ subjects $$\overline{G}_{ab}=\frac{1}{N}\underset{j=1}{\overset{N}{}}𝒢_{ab}^{(j)},\mathrm{where}a=U,L,b=U,L,$$ (5) and the ensemble total cross correlation $$\overline{C}_{UL}=\frac{\overline{G}_{UL}}{\left(\overline{G}_{LL}\overline{G}_{UU}\right)^{1/2}}.$$ (6) For the totality of the experimental data from all of the subjects we find $`\overline{C}_{UL}=0.92`$ almost completely anti-correlated. The above statistical analysis verifies objectively and quantitatively the view expressed in Sec.3. The hip strategy consists of a motion wherein the upper body and lower body randomly oscillate but in proportionately opposite directions. ## 5 Discussion In the study of condensed matter statistical mechanics, it is well known that one may tease out of the thermal equilibrium noise data many important mechanical parameters (Martin 1967). These parameters normally enter into the large dynamical displacements that would be present were those systems to be externally driven. Similarly, the notion has appeared in the literature that it would be possible to learn much about non-equilibrium motor biomechanics by merely observing the small displacement noise noise during quite standing (Lauk et al1998). This view certainly turns out to be true for the hip strategy employed to keep ones balance, as shown from the data reported here. Non-equilibrium hip motions are evident, for example, to those who have observed an infant in that time interval after knowing how to crawl but before learning how to walk (Shumway-Cook 1995). When not very quietly standing, the infant exhibits large hip motions while trying to stand. Since (during this critical period in life) the transition rate $`\mathrm{\Gamma }`$ for a fall (Koleva et al1999) obeys $`\mathrm{\Gamma }>0.05Hz`$, the infant often falls when the upper body coordinate is positive, either by sitting (backward fall) or returning to a crawl (forward fall). The non-equilibrium hip motions are equally evident in adults trying to balance on a tight rope with their forward direction normal to the tight rope direction. Since normal (non-circus trained) adults have a tight rope transition rate for a fall which also obeys $`\mathrm{\Gamma }>0.05Hz`$, the large hip strategy motions are clearly seen to exist somewhat before the inevitable fall onto a safety net. What has been shown in this work, is that the hip strategy motions are clearly in operation during a quite stance, even though the amplitude of hip motions are too small to observed by the unaided eye. With the aid of the ultrasound transducers in the SWA device, the two coordinate measurements clearly show an ongoing hip strategy in each and every subject tested. By a judicious choice of coordinates, other strategies should also yield to a quiet standing noise analysis technique. ## References
no-problem/9908/physics9908026.html
ar5iv
text
# Physical Parameters for Biconcave Shape Vesicles ## Equation and Requirements For an axisymmetric surface $`\mathrm{\Sigma }`$, we take the rotational axis to be $`z`$-axis and the plane of reflection $`xy`$-plane, then a biconcave shape is obtained by revolving and reflecting a curve, given by a function $`z(r)0`$ defined for $`r`$ in some interval $`[0,r_{\mathrm{}}]`$, where $`r_{\mathrm{}}`$ is the radius of the “equator”. For a biconcave axisymmetric surface, we expect $`z(r)`$ and its derivative $`w=z^{}`$ to have the following graphs. Usually, the Helfrich shape equation of axisymmetrc vesicles is written in term of the angle $`\psi `$ between the surface tangent and the plane perpendicular to rotational axis . For our discussion, it is convenient to rewrite it in terms of the derivative of the graph $`w(r)=z^{}(r)=\mathrm{tan}\psi `$. Then the Helfrich shape equation of axisymmetrc vesicles becomes an equation for $`w`$, which is $`{\displaystyle \frac{2rw^{\prime \prime }}{(1+w^2)^{5/2}}}`$ $`={\displaystyle \frac{5rww_{}^{}{}_{}{}^{2}}{(1+w^2)^{7/2}}}`$ (1) $`{\displaystyle \frac{2w^{}}{(1+w^2)^{5/2}}}+{\displaystyle \frac{2w+w^3}{r(1+w^2)^{3/2}}}`$ $`+{\displaystyle \frac{2c_0w^2}{1+w^2}}+{\displaystyle \frac{(c_0^2+\stackrel{~}{\lambda })rw}{(1+w^2)^{1/2}}}{\displaystyle \frac{\stackrel{~}{p}r^2}{2}}.`$ The biconcave shape and symmetries are then translated into suitable initial and boundary conditions, such as, $`w(0)=0`$; $`w^{}(0)=w_0^{}>0`$; $`w(r)\mathrm{}`$ as $`rr_{\mathrm{}}`$; an integral condition that $`\mathrm{}{\displaystyle _0^r_{\mathrm{}}}w\mathrm{d}r<0`$; and $`w`$ is required to have a unique local maximum and no other critical points. The lower order terms can be given by polynomials $`Q(t)`$ or its quadratic part $`R(t)=Q(t)t^3`$. After multiplying with $`rw^{}`$, the equation (1) can be written as follow, $$\begin{array}{cc}\hfill \left[\frac{r^2w_{}^{}{}_{}{}^{2}}{(1+w^2)^{5/2}}\right]^{}& =\left[\frac{w^2}{(1+w^2)^{1/2}}\right]^{}\hfill \\ & +r^3w^{}R(\kappa (r));\hfill \end{array}$$ (1a) $$\begin{array}{cc}\hfill \left[\frac{r^2w_{}^{}{}_{}{}^{2}}{(1+w^2)^{5/2}}\right]^{}& =\left[\frac{2}{\sqrt{1+w^2}}\right]^{}\hfill \\ & +r^3w^{}Q(\kappa (r)).\hfill \end{array}$$ (1b) where $`\kappa (r)={\displaystyle \frac{w}{r\sqrt{1+w^2}}}`$ is the principal curvature of $`\mathrm{\Sigma }`$ in the meridinal direction. Indeed, $$\kappa (r)w_0^{},\text{as }r0\text{,}$$ so the initial choice $`w_0^{}`$ is the central meridinal curvature or $`w_{0}^{}{}_{}{}^{2}`$ the Gaussian curvature. Note that if $`c_0=\stackrel{~}{\lambda }=\stackrel{~}{p}=0`$, equation (1a) can be integrated directly to a solution $`z=z(r)`$ which is a circular arc with radius $`1/w_0^{}`$. This is exactly the case when the meridinal curvature is a constant. ## Necessary Conditions The lower order terms in (1) capture the changes of the meridinal curvature. Since $`Q(0)=R(0)=\stackrel{~}{p}/2<0`$, if $`w_0^{}`$ is less than the smallest positive root of $`Q(t)`$, then both $`Q(w_0^{})<0`$ and $`R(w_0^{})<0`$. Our analysis shows that these negativity conditions are essential for the meridinal curvature $`\kappa (r)`$ to change sign. Furthermore, if the physical parameters constitute a polynomial $`Q`$ with only positive roots, a solution $`w(r)`$ with $`w_0^{}`$ small enough always satisfies all of our requirements. The threshold value of $`w_0^{}`$ can be calculated numerically if exact values of the parameters are given. The meridinal curvature $`\kappa (r)`$ of the vesicle $`\mathrm{\Sigma }`$ is the most basic quantity in our analysis. From equation (1b), if $`c_0>0`$ and the initial curvature at the center $`w_0^{}`$ is large, in the sense that $`R(w_0^{})>0`$, then $`\kappa (r)`$ increases and a biconcave shape cannot be formed. So a necessary condition for formation of axisymmetric solution of biconcave shape in the case that $`c_0>0`$ is $$R(w_0^{})=2c_0w_{0}^{}{}_{}{}^{2}+(c_0^2+\stackrel{~}{\lambda })w_0^{}\frac{\stackrel{~}{p}}{2}<0.$$ (2) The following pictures show the behaviors of $`\kappa (r)`$ for two different values of $`w_0^{}`$ with $`c_0=1`$, $`\stackrel{~}{\lambda }=0.25`$, and $`\stackrel{~}{p}=1`$. For these parameter values, $`R(0.277124)=0`$. $`R\left(w_0^{}\right)>0`$ $`R\left(w_0^{}\right)<0`$ The one on the left is taken with $`w_0^{}=0.278`$ so the necessary condition is not satisfied. In this case, $`\kappa `$ blows-up at finite radial distance and the solution only leads to a convex surface. The picture on the right is taken with $`w_0^{}=0.276`$. In this case, the necessary condition is satisfied, so $`\kappa `$ decreases and changes sign. But it only decays gradually and the integral area condition is not satisfied. As a result, the surface and its reflection intersects in $`^3`$. There is a more subtle boundary condition imposed by the reflection symmetry. One needs to verify that $`z(r)`$ and $`z(r)`$ together form a stationary solution with respect to any variation near the plane of reflection. For the situation that we discuss, it turns out that this is true. Moreover, we obtain a relation between the radius $`r_{\mathrm{}}`$ of the “equator” and the Gaussian curvature $`K(r_{\mathrm{}})`$ of the vesicle at any point on the “equator”, $$K(r_{\mathrm{}})^2=\frac{1}{r_{\mathrm{}}}Q\left(\frac{1}{r_{\mathrm{}}}\right).$$ (3) This relation is easily verified for the case that $`c_0=\stackrel{~}{\lambda }=\stackrel{~}{p}=0`$, in which, $`Q(t)=t^3`$ and $`K(r_{\mathrm{}})^2={\displaystyle \frac{1}{r_{\mathrm{}}^4}}`$. This is compatible with the fact that the solution surface is a round sphere. ## Sufficient condition To have a solution meeting all the requirements, we need to impose a stronger condition that $`w_0^{}`$ to be small enough, at least smaller than the positive roots of $`Q`$. This smallness assumption on $`w_0^{}`$ implies that not only $`R(w_0^{})<0`$ but also $`Q(w_0^{})<0`$. Both $`\kappa (r)`$ and $`w(r)`$ start initially positive and become negative when $`r`$ reaches a certain value $`r_0`$ (depending on the value of $`w_0^{}`$, actually, comparable to $`w_{0}^{}{}_{}{}^{1/2}`$ when $`w_0^{}`$ is small). In fact, we discover that, for any point of inflection $`r_c`$ of the graph of $`z(r)`$ with $`\kappa (r_c)>0`$, $$\frac{16}{3\stackrel{~}{p}}\underset{w_0^{}0}{lim}\frac{r_c^2}{w_0^{}}\underset{w_0^{}0}{lim}\frac{r_0^2}{w_0^{}}\frac{16}{\stackrel{~}{p}}.$$ As a consequence of these inequalities, the graph of $`z(r)`$ has a unique point of inflection. After $`r`$ reaches $`r_0`$, with the assumption that all roots of $`Q(t)`$ are positive, the lower order terms of equation (1b) are dominated and the function $`w(r)`$ blows down monotonically to $`\mathrm{}`$ in a finite radial distance. By further estimating the area according to $`w_0^{}`$, the integral area condition can be verified whenever $`w_0^{}`$ is sufficiently small. All these behaviors of the functions $`z(r)`$ and $`w(r)`$ with respect to $`Q`$ and the choice of $`w_0^{}`$ are illustrated in our summary on p. Discussions. The mathematical details will be presented in a forthcoming paper. ## Discussions Finally, we would like to remark on other solutions which do not correspond to biconcave surfaces. The one on the left is a solution for $`c_0=1`$, $`\stackrel{~}{\lambda }=3`$, $`\stackrel{~}{p}=2`$ with $`w_0^{}=2`$. Note that $`w_0^{}`$ is greater than the positive roots of $`R(t)`$ and $`Q(t)`$. Moreover, $`Q(t)`$ has negative roots and thus this is beyond the situation discussed above. In this case, after $`w=z^{}`$ becomes negative, it may not decrease monotonically to negative infinity, which is why a second “petal” may form. The other type is a rotational solution which is not symmetric under reflection in $`xy`$-plane . The picture is produced with $`c_0=1`$, $`\stackrel{~}{\lambda }=1`$ and $`\stackrel{~}{p}=1`$. The “upper” curve satisfies our equation (1) but not the area condition. The “lower” curve and the reflected “upper” curve are both solutions locally, so there is a bifurcation of solution at $`r_{\mathrm{}}`$. The surface only closes up by an asymmetric one. It should be remarked that, from our numerical study, not all solutions in the upper quadrant may be closed up by an asymmetric counterpart. For example, if the upper part satisfies the area condition, the only possible lower part is the reflection symmetric one. This suggests a possible uniqueness of biconcave solution. However, at the moment, there is not yet a mathematical proof. | Graph of $`Q`$ and positions of $`w_0^{}`$ ($`c_0=1`$, $`\stackrel{~}{\lambda }=0.25`$, $`\stackrel{~}{p}=1`$) | | | | --- | --- | --- | | Graphs of solution $`z`$ | | | | (a) $`w_0^{}=0.4`$ | (b) $`w_0^{}=0.25`$ | (c) $`w_0^{}=0.18`$ | | | | | | Graph of $`Q`$ and positions of $`w_0^{}`$ ($`c_0=1`$, $`\stackrel{~}{\lambda }=0.3`$, $`\stackrel{~}{p}=1`$) | | | | --- | --- | --- | | Graphs of solution $`z`$ | | | | (a) $`w_0^{}=0.4`$ | (b) $`w_0^{}=0.27`$ | (c) $`w_0^{}=0.15`$ | | | | | | Graph of $`Q`$ and positions of $`w_0^{}`$ ($`c_0=1`$, $`\stackrel{~}{\lambda }=0.25`$, $`\stackrel{~}{p}=1`$) | | | | --- | --- | --- | | Graphs of solution $`z`$ | | | | (a) $`w_0^{}=2`$ | (b) $`w_0^{}=0.5`$ | (c) $`w_0^{}=0.275`$ | | | | |
no-problem/9908/astro-ph9908016.html
ar5iv
text
# On the Nature of Soft X-ray Weak Quasi-Stellar Objects ## 1 Introduction While X-ray emission appears to be a universal property of Quasi-Stellar Objects (QSOs; Avni & Tananbaum 1986, hereafter AT86), there is evidence for a significant population of soft X-ray weak QSOs (SXW QSOs; Elvis 1992; Laor et al. 1997a; Yuan et al. 1998). These comprise $`10`$% of the optically selected QSO population and have soft X-ray emission that is $`\stackrel{>}{}`$10–30 times weaker than expected based on their continuum luminosities in other bands. In general, the cause of soft X-ray weakness has not been well established, and it could arise as a result of (1) intrinsic or intervening X-ray absorption, (2) an unusual underlying spectral energy distribution, or (3) extreme X-ray or optical variability. In this paper, we investigate the X-ray, ultraviolet and optical properties of a well-defined sample of SXW QSOs. Our main goals are (1) to determine if soft X-ray weakness correlates with other observables, (2) to examine the relationship between SXW QSOs, Broad Absorption Line QSOs (BAL QSOs) and QSOs with X-ray warm absorbers, and (3) to determine the reason for the low apparent soft X-ray luminosities of SXW QSOs. Regarding the second goal, it is known that BAL QSOs are generally weak in the soft X-ray band (e.g., Green & Mathur 1996), but the fraction of SXW QSOs that are BAL QSOs is poorly determined at present. Our focus is on objects that are more luminous and distant than local Seyferts (e.g., Maiolino & Rieke 1995). Our work differs from that of Green (1998) in that we study a smaller number of objects that are substantially more extreme in terms of their soft X-ray weakness. ## 2 Soft X-ray Weak QSO Selection We have selected our SXW QSOs from the Boroson & Green (1992; hereafter BG92) sample. This sample is composed of all 87 QSOs from the Bright Quasar Survey (BQS; Schmidt & Green 1983) with $`z<0.5`$, and our intention is to perform a thorough analysis of all the SXW QSOs in this sample. The BQS consists of the broad-line active galaxies with dominant star-like nuclei to the survey limits: $`B\stackrel{<}{}16.2`$ and $`UB\stackrel{<}{}0.44`$ (for Galactic latitudes greater than $`30^{}`$ and declinations above $`10^{}`$). To these limits it is thought to be $`\stackrel{>}{}70`$% complete at $`z<0.5`$ (see Véron et al. 1999), and a large amount of high-quality and uniformly analyzed data are available for the BG92 QSOs. Since the BQS QSOs were optically selected, the BQS should not be directly biased with respect to the inclusion or exclusion of SXW QSOs. Possible indirect biases are, unfortunately, inescapable when using current large-area QSO surveys (see Goodrich 1997 and Krolik & Voit 1998 for examples of potentially relevant indirect biases). Our selection method is based upon $`\alpha _{\mathrm{ox}}`$, the slope of a power law defined by the rest-frame flux densities at 3000 Å and 2 keV. That is, $`\alpha _{\mathrm{ox}}=0.372\mathrm{log}(f_{2\mathrm{keV}}/f_{3000\mathrm{\AA }})`$ where $`f_{2\mathrm{keV}}`$ and $`f_{3000\mathrm{\AA }}`$ are flux densities.<sup>1</sup><sup>1</sup>1Some authors prefer to work with $`\alpha _{\mathrm{ox}}^{}`$, the slope of a power law defined by the rest-frame flux densities at 2500 Å and 2 keV. For a 2500 Å to 3000 Å slope of $`\alpha _\mathrm{u}`$, we find $`\alpha _{\mathrm{ox}}^{}=1.03\alpha _{\mathrm{ox}}0.03\alpha _\mathrm{u}`$. Note that $`\alpha _{\mathrm{ox}}^{}`$, $`\alpha _{\mathrm{ox}}`$ and $`\alpha _\mathrm{u}`$ are negative numbers. While $`\alpha _{\mathrm{ox}}`$ calculations for many of the BQS QSOs have been performed before (e.g., Tananbaum et al. 1986), these are not sufficient for our purposes as they miss many of the SXW QSOs that we study here. These objects were either not observed or not detected. In addition, comparison of our $`\alpha _{\mathrm{ox}}`$ values with earlier ones allows us to search for $`\alpha _{\mathrm{ox}}`$ variability; extreme X-ray or optical variability could cause a QSO to appear soft X-ray weak (e.g., see Figure 1 of Boller et al. 1997). We have therefore calculated our own $`\alpha _{\mathrm{ox}}`$ values for the 87 BG92 QSOs following the method in the next paragraph. When possible we have determined the 3000 Å flux densities needed for $`\alpha _{\mathrm{ox}}`$ using the data from Neugebauer et al. (1987). For the 13 BG92 QSOs without such data, we calculated 3000 Å flux densities using data from De Bruyn & Sargent (1978; 4 QSOs), Laor et al. (1997ab; 2 QSOs), Neugebauer et al. (1979; 1 QSO) and Schmidt & Green (1983; 6 QSOs). We have used ROSAT data to determine the requisite 2-keV flux densities of our QSOs. We obtained count rates from the ROSAT All-Sky Survey (RASS; Voges et al. 1999) and pointed observations. All but four of our objects (0043+039, 1004+130, 1259+593, 1700+518) have ROSAT detections. Of these four, we note that 0043+039, 1004+130 and 1700+518 were also undetected in moderately deep ASCA and Einstein observations (Elvis & Fabbiano 1984; Gallagher et al. 1999). We converted the count rates into 2-keV flux densities using a power-law model with Galactic absorption. We used the Galactic column densities from Elvis, Lockman & Wilkes (1989), Lockman & Savage (1995) and Murphy et al. (1996). The power-law energy index ($`\alpha _\mathrm{x}`$) was predicted based on a fit to the H$`\beta `$ FWHM/$`\alpha _\mathrm{x}`$ correlation shown in Figure 5a of Laor et al. (1997a): $`\alpha _\mathrm{x}=6.1221.277\mathrm{log}(\mathrm{H}\beta \mathrm{FWHM})`$. The H$`\beta `$ FWHM/$`\alpha _\mathrm{x}`$ correlation has been directly established to hold for $`1/4`$ of the QSOs in our sample, and our fit is valid for H$`\beta `$ FWHM in the range $``$ 800–11000 km s<sup>-1</sup>. We excluded 3C 273 from our fit to prevent it from skewing the result due to its small error bar. The resulting $`\alpha _{\mathrm{ox}}`$ values are given in Table 1, and we show a histogram of these $`\alpha _{\mathrm{ox}}`$ values in Figure 1. For 0043+039 and 1004+130, we were able to place substantially tighter constraints upon $`\alpha _{\mathrm{ox}}`$ using the ASCA and Einstein results (see above). The $`\alpha _{\mathrm{ox}}`$ distribution is non-Gaussian with $`>99`$% confidence due to its tail towards large negative values of $`\alpha _{\mathrm{ox}}`$. We found acceptable agreement between our $`\alpha _{\mathrm{ox}}`$ values and those for the 23 QSOs given in Table 4 of Laor et al. (1997a). The average absolute difference in $`\alpha _{\mathrm{ox}}`$ for these 23 QSOs is 0.060, and no systematic discrepancies are apparent. We expect any systematic errors to be smaller than the random errors that are inherently present due to the X-ray and optical variability of QSOs. Furthermore, we have verified that none of the main results below depends upon the precise details of how we calculate $`\alpha _{\mathrm{ox}}`$. For example, we recover the same main results if we use $`\alpha _\mathrm{x}=1.72`$ (1.15) for all radio-quiet QSOs (radio-loud QSOs) in the $`\alpha _{\mathrm{ox}}`$ calculation above (see §4.1 of Laor et al. 1997a). In this paper we have defined a SXW QSO to be a QSO with $`\alpha _{\mathrm{ox}}2`$. These SXW QSOs thus have 2-keV flux densities that are $`\stackrel{>}{}25`$ times weaker than expected for a ‘normal’ QSO (see §4.3 of Laor et al. 1997a). While the precise choice of the $`\alpha _{\mathrm{ox}}`$ division value is somewhat arbitrary, inspection of the histogram shown in Figure 1 indicates that our choice is a reasonable one. The objects with $`\alpha _{\mathrm{ox}}2`$ stand out clearly from the bulk of the QSO population. We have tested for bimodality of our $`\alpha _{\mathrm{ox}}`$ distribution using the kmm mixture-modeling algorithm described by Ashman, Bird & Zepf (1994, hereafter ABZ94). This algorithm indicates bimodality of the distribution with high statistical significance (the derived kmm $`P`$-value is $`<0.02`$; see §2.1 of ABZ94). The algorithm places the 10 QSOs with the largest negative values of $`\alpha _{\mathrm{ox}}`$ into one group (corresponding to $`\alpha _{\mathrm{ox}}2`$) and the rest of the QSOs into another. The only QSO with an ambiguous kmm classification is 2251+113; this object is discussed further in §6. However, since the kmm algorithm makes several statistical assumptions (e.g., Gaussian-distributed populations; see Hartigan & Hartigan 1985 and ABZ94), we regard our bimodality result as suggestive rather than conclusive. With a sample of 87 QSOs, a distribution with a significant skew tail towards large negative $`\alpha _{\mathrm{ox}}`$ (compare §2 of AT86) could mimic a bimodal one. A nonparametric Hartigan & Hartigan (1985) dip test is unable to rigorously establish bimodality, but this certainly does not rule out bimodality since the dip test sacrifices a great deal of statistical power for its robustness. The histogram in Figure 1 may be compared with Figure 1 of AT86, although we note that our $`\alpha _{\mathrm{ox}}`$ values are substantially more complete and constraining. Our histogram may also be compared with Figure 10 of Yuan et al. (1998), although in this case our sample is more statistically complete and homogeneous (see §2 of AT86 for a discussion of the importance of complete samples). We find 10 SXW QSOs among the 87 QSOs in the BG92 sample, and these are listed in Table 2. Thus, SXW QSOs comprise $`11`$% of the optically selected QSO population. This percentage is in reasonable agreement with the ‘detections and bounds’ upper limit of AT86. All of our SXW QSOs are radio quiet (with $`\mathrm{log}R<1`$) except for 1004+130.<sup>2</sup><sup>2</sup>2Here and hereafter we take $`R`$ to be the ratio of the observed 6 cm and 4400 Å flux densities. In Table 2 we have also given the historical values for $`\alpha _{\mathrm{ox}}`$ from Einstein observations (Tananbaum et al. 1986); these values will be used in §8. It is possible that 1259+593 is a SXW QSO as well, but unfortunately the current X-ray data do not allow a tight upper limit to be placed upon $`\alpha _{\mathrm{ox}}`$ for this object ($`\alpha _{\mathrm{ox}}<1.75`$). We also comment that 0844+349 was seen to be fairly weak in soft X-rays during a ROSAT pointed observation but not during the RASS (Yuan et al. 1998). Here we have adopted the RASS count rate, and we discuss this object further in §6. ## 3 Statistical Analyses of Continuum and Optical Emission-Line Properties We have performed statistical tests to compare the continuum and optical emission-line properties of SXW QSOs with those of the bulk of the QSO population. In particular, we have compared the properties of our 10 SXW QSOs (hereafter SXW for this group) to those of the other 77 BG92 QSOs (hereafter non-SXW for this group) using two-sample Kolmogorov-Smirnov (hereafter KS) and Kuiper tests (see Press et al. 1992). We chose to apply both of these tests because they statistically complement each other. These tests were objectively chosen in an a priori manner; we did not inspect the data in detail and then decide which statistical tests to apply. The main properties we compare are those listed in Tables 1 and 2 of BG92.<sup>3</sup><sup>3</sup>3For 1211+143 we have corrected the $`R`$ parameter following Kellermann et al. (1994). We have also corrected the H$`\beta `$ FWHM of 1307+085 to be 5320 km s<sup>-1</sup> and the H$`\beta `$ shift of 1351+236 to be 1.099. In addition, we compare the degree of broad-band optical continuum polarization (3200–8600 Å; given in Table 1 of Berriman et al. 1990), the optical continuum slope ($`\alpha _{\mathrm{opt}}`$; given in Table 5 of Neugebauer et al. 1987), and the radio structure (e.g., Miller, Rawlings & Saunders 1993; Kellermann et al. 1994). The probabilities resulting from our KS and Kuiper tests are listed in Table 3, and some of the continuum and optical emission-line properties are illustrated in Figure 2. ### 3.1 Similarities In terms of many of their properties, SXW QSOs are homogeneously distributed among the QSO population. For example, there are no statistically significant differences between SXW and non-SXW for $`M_V`$, $`z`$, or $`R`$. Of the 71 BG92 radio-quiet QSOs, 9 are soft X-ray weak. Of the 16 BG92 radio-loud QSOs, 1 is soft X-ray weak (the radio-loud QSO 2251+113 is also quite weak in the soft X-ray band; see §6). These rates of occurrence are statistically consistent according to a Fisher exact probability test for a $`2\times 2`$ contingency table. In addition, the equivalent width (EW) distributions of H$`\beta `$, He ii and Fe ii are consistent between SXW and non-SXW. As a class, our SXW QSOs do not show higher broad-band optical continuum polarization than the non-SXW QSOs. Of our 10 SXW QSOs, 9 have broad-band polarization percentages smaller than 0.9%. The one notable exception is 1535+547 which showed the highest polarization percentage in the BQS: 2.5% (see Berriman et al. 1990). We give additional notes on the polarization properties of our SXW QSOs in §5. We also find no evidence for a statistically significant difference between the optical continuum slopes of SXW and non-SXW (i.e. the SXW QSOs do not appear to be systematically redder than the other BG92 QSOs). Although we note that both SXW and non-SXW were drawn from the BQS which employed selection by ultraviolet excess, the range in ultraviolet excess for the BQS is sufficiently large that a real difference would not necessarily be masked. ### 3.2 Differences There are several ways in which the optical emission-line properties of SXW QSOs stand out from those of the bulk of the optically selected QSO population. The most notable one is that SXW QSOs systematically have low \[O iii\] luminosities, and we have further illustrated this in Figure 3. In addition, while the H$`\beta `$-FWHM and H$`\beta `$-asymmetry distributions of SXW and non-SXW are consistent, they have significantly different H$`\beta `$-shift and H$`\beta `$-shape distributions. The H$`\beta `$-shape distributions of SXW and non-SXW are found to be inconsistent by both the KS and Kuiper tests, and SXW QSOs preferentially have ‘peaky’ line profiles with concave sides. The H$`\beta `$-shift distributions of SXW and non-SXW are only shown to be different with high statistical significance by the Kuiper test. This result can be understood by inspection of §14.3 of Press et al. (1992) and Figure 2. The SXW QSOs tend to lie at either the redward or blueward extremes of the H$`\beta `$-shift parameter, and this is where the KS test is least sensitive but the Kuiper test is significantly more sensitive (i.e. the KS test is good at finding differences in the median value but not as good at finding differences in the ‘spread’). While we must recommend some caution regarding the H$`\beta `$-shift result since it is based on only 8 SXW data points (1700+518 and 2112+059 do not have H$`\beta `$-shift measurements in BG92 due to their weak \[O iii\] emission), it does appear to be reasonably robust. If we add a hypothetical SXW data point with an H$`\beta `$ shift of 0, we still find a significant difference between the H$`\beta `$-shift distributions of SXW and non-SXW. In this case, the Kuiper test probability is 0.0132. SXW QSOs tend to lie toward the weak-\[O iii\] end of ‘eigenvector 1’ as defined by BG92. Eigenvector 1 describes a set of optical emission-line properties that vary together in a highly coordinated manner (e.g., \[O iii\] strength, Fe ii strength, H$`\beta `$ FWHM, and H$`\beta `$ asymmetry), and it emerged from the Principal Component Analysis (PCA) performed by BG92. Note, however, that eigenvector 1 was derived from a PCA that included $`\alpha _{\mathrm{ox}}`$ (see Table 4 of BG92). In order to be sure that the weak dependence of eigenvector 1 upon $`\alpha _{\mathrm{ox}}`$ is not affecting our result, we have performed our own PCA with $`\alpha _{\mathrm{ox}}`$ excluded. We find that our SXW QSOs still tend to lie at the weak-\[O iii\] end of this ‘$`\alpha _{\mathrm{ox}}`$-free’ eigenvector 1 (see Table 3). ### 3.3 Radio Structures The radio structures of SXW and non-SXW are more difficult to compare in a rigorous statistical manner due to the difficulty involved in quantifying radio structure. Our one radio-loud SXW QSO, 1004+130, is lobe dominated, suggesting that we are viewing the active nucleus in a relatively edge-on manner (see Wills, Brandt & Laor 1999 for detailed discussion). There are no obvious radio structure differences between the radio-quiet members of SXW and non-SXW, although we note that better radio imaging data are needed to properly examine this matter. The only radio-quiet SXW QSO with radio structure of particular note is 1700+518. It shows a double source morphology (e.g., see §5.1.3 of Miller et al. 1993 and §3.1 of Kukula et al. 1998). 2251+113 is radio-loud and fairly soft X-ray weak (see §6), and it also is lobe dominated. ## 4 Ultraviolet Absorption-Line Properties of Soft X-ray Weak QSOs If photoelectric X-ray absorption is the primary cause of soft X-ray weakness for QSOs, then the X-ray absorbing material might also be detected at other wavelengths. The obvious place to look is in the ultraviolet where cosmically abundant elements have strong bound-bound transitions (e.g., Mathur et al. 1994). Alternatively, X-ray and ultraviolet absorption might have a high frequency of joint occurrence even if the two types of absorption do not arise in the same gas (e.g., Crenshaw et al. 1998 and references therein). We have therefore collected the available HST and IUE spectra for the BG92 QSOs to examine if SXW QSOs preferentially show strong ultraviolet absorption. We have focused on absorption in the C iv $`\lambda 1549`$ line (hereafter C iv) because the most data are available for this region. There is a wide range in the quality of the spectra, and spectra with particularly poor signal to noise were rejected (see below). Preference was usually given to HST data. We found 46 objects with adequate HST coverage of at least the C iv region and 9 additional objects with adequate IUE coverage. When several datasets were available we co-added them, applying any needed shifts in wavelength scale and weighting appropriately. The rest wavelength scale was determined from a redshift based on the best \[O iii$`\lambda 5007`$ data where possible. We have searched all the resulting spectra for BALs or other associated C iv absorption features, and in Table 1 we give our measured rest-frame EWs. All measurements were performed by one author with substantial experience in HST and IUE data analysis (BJW), and particular effort was made to ensure consistency and uniformity. The measurements were made independently of knowledge of $`\alpha _{\mathrm{ox}}`$. We excluded narrow Galactic interstellar absorption lines based on line identifications over the available spectral range. We were able to confirm some weak C iv absorption features from the doublet wavelength separations or the presence of corresponding absorption features (e.g., N v $`\lambda 1240`$, Ly$`\alpha `$, O vi $`\lambda 1034`$) in the same redshift system. Generally, only measurements with signal to noise greater than 4 were retained. When no absorption was detected we enter 0 for the EW. While our EWs refer to measurable features, they are probably also a reasonable measure of the total EW of C iv associated absorption for a given object. However, we do not give detailed uncertainties. This is because it is difficult to give meaningful limits to any unmeasured associated absorption. Such limits are a complicated function of many parameters: noise, the unknown widths of possible absorption features, the effective continuum placement (especially in the presence of broad emission line wings), and the actual level of the local continuum (which may be low in the true continuum to high near the center of a broad emission line). The archival data are readily accessible to the interested reader, and some examples of spectra are given in §5 and §6. In Figure 4 we show a plot of $`\alpha _{\mathrm{ox}}`$ versus C iv absorption EW. This plot reveals a striking connection between soft X-ray weakness and strong C iv absorption. Note that the ordinate of this plot spans a range of $`\stackrel{>}{}800`$ in C iv absorption EW. We have labeled our SXW QSOs and other QSOs of particular interest, and we give object-specific notes on these QSOs in §5 and §6. At least eight of our ten SXW QSOs show C iv absorption with EW $`>1`$ Å, and most have EW $`>5`$ Å. We have performed a Spearman rank-order correlation analysis using the data shown in Figure 4 (including appropriate censoring; Isobe, Feigelson & Nelson 1986) and find the correlation to be significant with $`>99`$% confidence. We have searched for other correlations that could be inducing the correlation of Figure 4 as a secondary effect, and we find none. For example, we do not find any correlation between H$`\beta `$ FWHM and C iv absorption EW. This correlation is worth checking because $`\alpha _\mathrm{x}`$ depends upon H$`\beta `$ FWHM (see §2); this relationship could have induced a weak dependence of $`\alpha _{\mathrm{ox}}`$ on H$`\beta `$ FWHM. In Figure 4 we have also marked bona-fide BAL QSOs with asterisks.<sup>4</sup><sup>4</sup>4To be spoken of as a bona-fide BAL QSO, we require a QSO to satisfy the non-zero balnicity requirement outlined in §3.1 of Weymann et al. (1991). To our knowledge, the only BG92 QSOs that clearly satisfy this requirement are 0043+039, 1001+054, 1700+518, 2112+059 and probably 1004+130 (see §5). These 5 BAL QSOs have the largest C iv absorption EW values in our SXW QSO sample, and inspection of Table 2 reveals that they are also our 5 most optically luminous SXW QSOs. In light of this result, we have used a Spearman analysis to search for a correlation between $`M_\mathrm{V}`$ and C iv absorption EW for our 10 SXW QSOs, and we find a correlation that is significant with $`>99`$% confidence. However, with only 8 data points and 2 upper limits we cannot rigorously demonstrate a causative relation between $`M_\mathrm{V}`$ and C iv absorption EW for our SXW QSOs. Finally, we have used a Spearman analysis to search for a correlation between the C iv absorption EW and the broad-band optical continuum polarization (again from Berriman et al. 1990) for our 55 BG92 QSOs with C iv data. The analysis indicates a correlation to be present at the $`>96`$% confidence level, but a plot of these two quantities shows that any correlation has a large scatter. ## 5 Notes on Individual Soft X-ray Weak QSOs In this section we give relevant comments on some of our SXW QSOs. In Figure 5 we show spectra of the C iv region for our SXW QSOs, and the reader should refer to this figure while reading this section. 0043+039 and 1700+518 are well-known BAL QSOs that have been intensively studied, so we shall not give comments on them here. 1001+054: This QSO was first noted to be soft X-ray weak by Laor et al. (1997), and it has a narrow-line type 1 spectrum. HST spectra reveal broad C iv and Ly$`\alpha `$ absorption, and we measure a small, but non-zero, balnicity index (see Weymann et al. 1991) of $`420`$ km s<sup>-1</sup>. A detailed study of the HST data will be presented in Wills et al., in preparation. 1004+130 (PKS 1004+130, 4C 13.41): This is the only radio-loud SXW QSO in our sample ($`\mathrm{log}R=2.36`$ and the radio spectrum is steep). Its polarization percentage rises with frequency, reaching $`2`$% by 4000 Å (Antonucci et al. 1996 and references therein). It was first noted to be soft X-ray weak by Elvis & Fabbiano (1984), who studied it using data from Einstein and IUE. In the low-resolution IUE spectra Kinney et al. (1991) recognized unresolved N v and C iv absorption near the emission-line redshift. A high-resolution HST GHRS spectrum was obtained by Bowen et al. (1997) to search for intervening absorption by the low-redshift dwarf spheroidal galaxy Leo I. Serendipitously, this spectrum showed the associated N v absorption doublet with at least two velocity components (Figure 3 of Bowen et al. 1997). In this same spectrum, which we retrieved from the HST archives, we discovered the corresponding high-ionization absorption lines of the O vi doublet (see Figure 6). These appear to be optically thick and suggest the presence of partial covering. In addition, our detailed analyses of the IUE and HST data suggest much broader unresolved troughs of C iv, Si iv and O vi, extending to outflow velocities of $`\mathrm{10\hspace{0.17em}000}`$ km s<sup>-1</sup>. We estimate a balnicity index of $`850`$ km s<sup>-1</sup> and believe that 1004+130 is probably a low-redshift, radio-loud BAL QSO. There is probable Ly$`\alpha `$ absorption and strong N v absorption too. This absorption affects the Ly$`\alpha `$ emission line and accounts for its weakness in this QSO. A detailed analysis of the ultraviolet data is presented by Wills, Brandt & Laor (1999). 1011–040: The archival IUE data for this SXW QSO do not reveal any obvious associated absorption, but the spectral quality is not high. It seems unlikely but not impossible that the blue wing of the C iv emission line could be stronger than observed, implying a larger absorption EW. In this case the observed emission line could appear redshifted. However, the peaks of the Ly$`\alpha `$ and C iv emission lines give a consistent redshift, and the symmetry of the stronger Ly$`\alpha `$ line does not hint at significant absorption. 1126–041 (Mrk 1298): Wang et al. (1999a) have recently studied the ultraviolet and X-ray absorption in the spectrum of this QSO. HST spectra of this object would allow the strong ultraviolet absorption to be studied in significantly more detail. 1411+442: 1411+442 was noted to be soft X-ray weak by Laor et al. (1997), and its ultraviolet absorption and other properties are discussed in detail by Malkan, Green & Hutchings (1987) and Wang et al. (1999b). While its ultraviolet absorption is prominent, it is not a bona-fide BAL QSO. Brinkmann et al. (1999) have presented the ASCA data for this QSO, and our independent analysis of these data is in general agreement with theirs. The data suggest absorption by an intrinsic column density of $``$ (1–2)$`\times 10^{23}`$ cm<sup>-2</sup>, and there is also evidence for a scattered X-ray component at low energies. The ASCA spectrum has limited photon statistics, and the X-ray column density is subject to significant uncertainty because the underlying continuum shape and the ionization state of the X-ray absorber cannot be tightly constrained. 1535+547 (Mrk 486): This object is our weakest detected SXW QSO, and it is also the most highly polarized BQS QSO. Its polarization (2–8%) and associated C iv absorption have recently been studied by Smith et al. (1997), and they argue for a complex broad emission-line region that contains dust. 1535+547 has a narrow-line type 1 spectrum. 2112+059: The HST spectrum for this SXW QSO reveals broad ultraviolet absorption by C iv and Ly$`\alpha `$ (Jannuzi et al. 1998). The broad absorption tends to be shallow, but there are narrower absorption components as well. We measure a balnicity index of $`2980`$ km s<sup>-1</sup>. 2112+059 also has an intervening damped Ly$`\alpha `$ absorption system at $`z=0.2039`$ (e.g., Lanzetta, Wolfe & Turnshek 1995), but this system does not materially affect our analysis of the associated ultraviolet absorption. We do not expect this system to cause substantial neutral or ionized X-ray absorption; its neutral hydrogen column density is $`2.5\times 10^{20}`$ cm<sup>-2</sup>, and we expect any ionized column density to be $`\stackrel{<}{}5\times 10^{19}`$ cm<sup>-2</sup>. 2214+139 (Mrk 304): The archival IUE spectrum for this SXW QSO does not reveal any obvious associated absorption, but the spectral quality is not high (also see Clavel & Joly 1984). It is not impossible that the blue wing of the C iv emission line could be stronger than it appears, hiding significant absorption. However, the stronger, symmetric Ly$`\alpha `$ emission line shows no hint of corresponding blueshifted absorption. Rachen et al. (1996) noted that the flatness of the ROSAT PSPC spectrum might be due to photoelectric absorption, but detailed spectral analysis of these data is not possible due to the limited photon statistics. ## 6 Notes on Other Relevant QSOs Here we give comments on other objects that lie in interesting locations in Figure 4. The reader should refer to this figure while reading this section. 0844+349: Yuan et al. (1998) found 0844+349 to be a SXW QSO during a pointed ROSAT observation but not during the RASS. During the pointed observation they found $`\alpha _{\mathrm{ox}}^{}=2.05`$ (calculated between 2500 Å and 2 keV). For these data we calculate $`\alpha _{\mathrm{ox}}=1.83`$ using the methods in §2, and thus 0844+349 does not quite satisfy our criterion for a SXW QSO. This QSO was also not soft X-ray weak during an Einstein IPC observation (Kriss 1988). The current data suggest that 0844+349 is not usually a SXW QSO, and it appears to have been caught in an unusually low flux state during the pointed ROSAT observation. It has shown soft X-ray variability by a factor of $`6`$, and this may be due to time-variable X-ray absorption. The poor statistics of the pointed ROSAT spectrum make it difficult to examine this issue in detail, especially if significant X-ray scattering is present. Corbin & Boroson (1996) comment that this QSO shows ‘associated absorption’ in an HST spectrum, and our analysis of the HST data they used reveals narrow associated Ly$`\alpha `$ absorption. There are no HST data for the C iv region, but co-added IUE data suggest C iv absorption consistent in velocity with the Ly$`\alpha `$ absorption. 1114+445: This QSO shows moderate strength X-ray and ultraviolet absorption by ionized gas (George et al. 1997; Mathur, Wilkes & Elvis 1998). Its intermediate position relative to the low-absorption QSOs and SXW QSOs in Figure 4 is notable. 1114+445 is one of the more highly polarized BQS QSOs (2.3%; Berriman et al. 1990). 1259+593: As noted in §2, this object has only a poor limit on its X-ray flux, and we cannot at present determine if it satisfies our $`\alpha _{\mathrm{ox}}2`$ criterion to be a SXW QSO. Bahcall et al. (1993) presented an HST spectrum for this object, and there is no obvious intrinsic absorption (also see Lu & Zuo 1994). We do not expect significant X-ray absorption from high-velocity cloud Complex $`C`$ since its neutral hydrogen column density in this direction is $`1.5\times 10^{20}`$ cm<sup>-2</sup> (see Savage et al. 1993 and references therein). Similarly, UGC 8040 and UGC 8046 should not cause significant X-ray absorption given the ultraviolet constraints on neutral hydrogen absorption lines and the possible ionization levels in the interstellar media of these galaxies (see Bowen et al. 1996 and references therein). 1309+355: This is a flat-spectrum radio-intermediate QSO ($`\mathrm{log}R=1.26`$), and Miller et al. (1993) and Falcke et al. (1996) argue that it is a boosted radio-quiet QSO. The radio properties may indicate that the central engine is viewed in a relatively pole-on manner. In light of this, the strong ultraviolet absorption lines seen in the HST spectrum (C iv, N v, O vi and Ly$`\alpha `$) are notable. This object has $`\alpha _{\mathrm{ox}}=1.71`$; it is moderately X-ray weak, especially for a radio-loud QSO. 1351+593: This QSO is moderately weak in soft X-rays (also see Elvis 1992), and it shows associated ultraviolet absorption in HST, HUT and IUE spectra (Brosch & Gondhalekar 1984; Granados et al. 1993; Zheng et al. 1999). The absorption runs to fairly high velocities ($`3000`$ km s<sup>-1</sup>), but 1351+593 probably does not qualify as a bona-fide BAL QSO. 1402+261: This QSO shows moderate strength ultraviolet absorption by Ly$`\alpha `$ and C iv. The available ROSAT data are consistent with no intrinsic X-ray absorption by either neutral or ionized gas (Ulrich-Demoulin & Molendi 1996; Laor et al. 1997). 1404+226: 1404+226 has one of the most extreme narrow-line type 1 spectra for a QSO, and its X-ray spectrum shows a strong soft X-ray excess as well as additional poorly understood complexity (e.g., Ulrich-Demoulin & Molendi 1996; Leighly et al. 1997; Ulrich et al. 1999). The additional complexity may be associated with X-ray absorption edges or lines. HST spectra for 1404+226 reveal moderate strength ultraviolet absorption (Ulrich et al. 1999). 1425+267: Laor et al. (1997) pointed out that 1425+267 was moderately soft X-ray weak compared to other radio-loud QSOs, and it has a clear double-lobed radio structure. Complex residuals below 1 keV in archival ASCA spectra suggest the presence of intrinsic X-ray absorption, although the absorption cannot be probed in detail due to limited photon statistics. The HST spectrum for 1425+267 shows moderate-strength absorption by Ly$`\alpha `$ and C iv. 1552+085: Turnshek et al. (1997) suggest that this QSO has a C iv BAL in its IUE spectrum, but the data are noisy and better spectra are needed to reliably study ultraviolet absorption. For this reason, we do not quote a C iv absorption EW in Table 1, and this object does not appear in Figure 4. We find an $`\alpha _{\mathrm{ox}}`$ value of $`1.77`$, which is fairly small but still somewhat larger than is typically seen for BAL QSOs. 1552+085 is one of the more highly polarized BQS QSOs (1.9%; Berriman et al. 1990). 1704+608 (3C351): This is a lobe-dominated, radio-loud QSO ($`\mathrm{log}R=2.81`$) with moderate strength X-ray and ultraviolet absorption by ionized gas (e.g., Mathur et al. 1994; Nicastro et al. 1999). Like 1114+445 and 1425+267, it occupies an intermediate position between low-absorption QSOs and our SXW QSOs (also see Fiore et al. 1994). 1704+608 does not have notable polarization. 2251+113: This steep-spectrum radio-loud QSO ($`\mathrm{log}R=2.56`$) is moderately weak in the soft X-ray band, and associated ultraviolet absorption by Ly$`\alpha `$, N v, C iv and Si iv has been detected by HST (Bahcall et al. 1993). We have analyzed an archival ASCA spectrum of this QSO, and these data suggest absorption by an intrinsic column density of $`\stackrel{>}{}3\times 10^{21}`$ cm<sup>-2</sup>. However, as for 1411+442 and 1425+267, the ASCA spectrum has limited photon statistics and hence the X-ray column density is subject to significant uncertainty. ## 7 $`𝜶_{\mathrm{𝐨𝐱}}`$ Correlations after Removal of the Soft X-ray Weak QSOs We have used our improved $`\alpha _{\mathrm{ox}}`$ values for the BG92 QSOs to examine correlations between $`\alpha _{\mathrm{ox}}`$ and other observables in the most general manner possible. In these correlation analyses, we have excluded our 10 SXW QSOs from consideration. This is important since here we are interested in physical correlations that are not affected by objects where $`\alpha _{\mathrm{ox}}`$ has been altered by strong X-ray absorption (see §8 for the evidence that soft X-ray weakness is primarily due to absorption). We have performed Spearman rank-order correlations between our $`\alpha _{\mathrm{ox}}`$ values and the quantities tabulated by BG92, and our results are given in Table 4. In agreement with Green (1998), we find significant ($`P_{\mathrm{Spearman}}<0.04`$) correlations between $`\alpha _{\mathrm{ox}}`$ and the EWs of H$`\beta `$, \[O iii\], He ii and Fe ii. In addition, we find the $`\alpha _{\mathrm{ox}}`$ correlations against He ii/H$`\beta `$, Fe ii/H$`\beta `$, H$`\beta `$ FWHM, H$`\beta `$ asymmetry, $`\alpha _{\mathrm{opt}}`$, eigenvector 1, eigenvector 2 (see §3.2 of BG92 for discussion), and the ‘$`\alpha _{\mathrm{ox}}`$-free’ eigenvector 1 to be significant. The absence of a significant correlation with $`R`$ is notable. Earlier studies have found a correlation between $`\alpha _{\mathrm{ox}}`$ and $`R`$ (e.g., Zamorani et al. 1981, hereafter Z81), and this apparent discrepancy is discussed below in §8.2. Our correlations supersede and extend those of BG92 since (1) our $`\alpha _{\mathrm{ox}}`$ values are more complete and constraining than those used by BG92 and (2) we have removed objects with evidence for strong X-ray absorption (see above). ## 8 Discussion and Conclusions ### 8.1 The Origin and Demography of Soft X-ray Weak QSOs We have systematically investigated the nature of the SXW QSOs in the $`z<0.5`$ BQS. On the whole, the data support the idea that X-ray absorption is the primary cause of soft X-ray weakness in QSOs. We detect remarkably strong (EW $`>4.5`$ Å) C iv absorption in 8 of our 10 SXW QSOs, and ultraviolet and X-ray absorption have a high probability of joint occurrence. For comparison, only 1 of 45 non-SXW QSOs with C iv coverage shows absorption with EW $`>4.5`$ Å. The two SXW QSOs without clear ultraviolet absorption, 1011–040 and 2214+139, both have IUE spectra of only limited quality, and HST spectra are needed to further examine the possibility of ultraviolet absorption. ASCA data are available for only three of our SXW QSOs: 0043+039, 1411+442 and 1700+518. The SXW QSOs 0043+039 and 1700+518 were not detected (consistent with the presence of heavy absorption; see Gallagher et al. 1999), and 1411+442 shows evidence for substantial absorption in its X-ray spectrum (see §5). Furthermore, we have identified a general correlation between $`\alpha _{\mathrm{ox}}`$ and C iv absorption EW that appears to be due to a continuum of absorption properties (see Figure 4). Unabsorbed QSOs and SXW QSOs lie towards opposite ends of this correlation, while QSOs with X-ray warm absorbers lie near the middle of the correlation (see the extensive object notes in §5 and §6). BAL QSOs, which comprise a subclass of SXW QSOs, appear to lie at the absorbed extreme of the correlation; the bona-fide BAL QSOs in our SXW QSO sample have ultraviolet absorption that is the strongest we observe, and they often have only upper limits on their X-ray fluxes. The observed correlation is generally consistent with models that postulate a connection between orientation and absorption strength. For example, as one increases the inclination angle one might move through the sequence: unabsorbed QSO, X-ray warm absorber QSO, non-BAL SXW QSO, BAL SXW QSO, and perhaps type 2 QSO. However, our results in §8.2 regarding \[O iii\] luminosity suggest that this picture cannot be complete and that an intrinsic property must also play an important role. In addition, we note that there is evidence suggesting the central engine in the moderately soft X-ray weak QSO 1309+355 is viewed in a relatively pole-on manner (see §6). Finally, the potential relation between $`M_\mathrm{V}`$ and C iv absorption EW for our SXW QSOs (see §4) would be difficult to understand in the context of a pure orientation model. Although the correlation in Figure 4 strongly suggests that soft X-ray weakness is due to absorption, and that the X-ray and ultraviolet absorbers are related, this correlation does not imply that these absorbers are identical. In fact, a uniform screen which completely covers both the X-ray and ultraviolet emission sources is not expected to produce the shape of the correlation seen in Figure 4. This follows since the ultraviolet absorption is by a resonance line, where the absorption EW grows with the absorbing column $`N_\mathrm{H}`$ following a ‘curve of growth’ (e.g., Chapter 14 of Gray 1992), while the X-ray absorption is by bound-free edges, and thus grows like $`e^{\tau _{\mathrm{bf}}}`$ where $`\tau _{\mathrm{bf}}N_\mathrm{H}`$. One expects $`\alpha _{\mathrm{ox}}`$ to be practically independent of the C iv EW up to an $`N_\mathrm{H}`$ which gives $`\tau _{\mathrm{bf}}1`$, and above this one expects a very rapid drop in $`\alpha _{\mathrm{ox}}`$ associated with a very slow increase in the C iv EW. Instead, Figure 4 indicates a gradual increase in the C iv EW is associated with a gradual decrease in $`\alpha _{\mathrm{ox}}`$. This may be explained if both the X-ray and ultraviolet absorption are optically thick, but with an absorption covering factor which is less than unity (possibly due to scattering). In this case both $`\alpha _{\mathrm{ox}}`$ and the C iv EW just provide an indication of the X-ray and ultraviolet absorber covering factors, which may vary together. Other conceivable causes of soft X-ray weakness include (1) an unusual underlying spectral energy distribution and (2) extreme X-ray or optical variability. However, we consider it unlikely that either of these is the primary cause of soft X-ray weakness in optically selected QSOs. Regarding possibility 1, SXW QSOs might be postulated to be intrinsically weak emitters of soft X-rays. Our optical line studies described in §3 allow us to address this since line EWs depend upon the shape of the ionizing spectral energy distribution. We do not, for example, find evidence for a statistically significant difference in the He ii EW distributions for SXW QSOs and non-SXW QSOs. The creation of the He ii line is driven by 54–150 eV continuum photons, and it would thus be difficult to produce a normal strength line if the underlying spectral energy distribution were anomalously weak in this energy range (see Korista, Ferland & Baldwin 1997 and references therein). Regarding possibility 2 of the previous paragraph, extreme X-ray variability (with variability amplitude $`\stackrel{>}{}10`$) has been seen for some BQS QSOs, while such extreme optical variability appears much rarer (e.g., Giveon et al. 1999). To investigate potential $`\alpha _{\mathrm{ox}}`$ variability among our 10 SXW QSO, we have compared our $`\alpha _{\mathrm{ox}}`$ measurements and upper limits with the 3 $`\alpha _{\mathrm{ox}}`$ measurements and 3 $`\alpha _{\mathrm{ox}}`$ upper limits of Tananbaum et al. (1986; see Table 2). While the data are limited, we do not find evidence for outstanding $`\alpha _{\mathrm{ox}}`$ variability. Our results imply that selection by soft X-ray weakness is an effective ($`\stackrel{>}{}80`$% successful) way to find low-redshift QSOs with strong ultraviolet absorption. This is important from a practical point of view because the optical and X-ray flux densities needed to establish soft X-ray weakness can be obtained in an inexpensive manner from publicly available optical images (e.g., the Palomar Schmidt and UK Schmidt sky surveys) and X-ray data (e.g., the RASS and ROSAT pointed observations). When data from the RASS are used, we expect this method to be effective down to about $`B=17`$; for (unabsorbed) QSOs with lower optical fluxes, the expected X-ray flux becomes comparable to or less than the RASS sensitivity limit. This selection method appears to be significantly more effective than others that have been developed to find QSOs with strong ultraviolet absorption (e.g., Turnshek et al. 1997). We have found the percentage of SXW QSOs in the optically selected QSO population (specifically, the $`z<0.5`$ BQS) to be $`11`$%. This percentage is in general agreement with previous rough estimates but is more statistically reliable since it is derived from a larger and better-defined sample. We find 4–5 bona-fide BAL QSOs (0043+039, 1001+054, 1700+518, 2112+059 and probably 1004+130) in our sample, and if all BAL QSOs are soft X-ray weak then our methods should have identified all the BAL QSOs in the $`z<0.5`$ BQS. It is of interest to examine if the BAL incidence we find is consistent with that found in other QSO surveys (see §11.1 of Krolik 1999). Weymann (1997) gives a BAL QSO incidence of about $`11`$% for the $`z=`$ 1.4–3.0 QSOs from the Large Bright Quasar Survey (LBQS; see Hewett, Foltz & Chaffee 1995), and we would have expected to detect $`9.6`$ BAL QSOs given this percentage.<sup>5</sup><sup>5</sup>5We believe the ‘true’ 11% incidence rather than the ‘raw’ 8% incidence is the appropriate number to use for comparison in this case (see Weymann 1997). At the wavelengths at which they were selected, the $`z<0.5`$ BQS QSOs do not have their fluxes diminished by BAL troughs. The number of BALQSOs we detect, while perhaps somewhat smaller, does not show any strong inconsistency with the expectation from the LBQS. A Fisher exact probability test for a $`2\times 2`$ contingency table gives a probability of $`5`$% when we compare the BQS and LBQS samples. Our results are in general agreement with the idea that the incidence of BALs in modest-redshift QSOs is about the same as in higher-redshift QSOs. As mentioned in §4, the 5 bona-fide BAL QSOs in our SXW QSO sample are our 5 most optically luminous SXW QSOs. They also have the largest C iv absorption EWs, and among our SXW QSOs we find suggestive evidence for a correlation between $`M_\mathrm{V}`$ and C iv absorption EW. If the outflows from our SXW QSOs are optically thick, the C iv absorption EW is expected to be proportional to just the outflow velocity spread and the outflow covering factor, and the outflow velocity spread is plausibly related to the ultraviolet luminosity (e.g., see equation 6 of Scoville & Norman 1995). Thus a relation between $`M_\mathrm{V}`$ and C iv absorption EW seems physically plausible, but further work is clearly needed to examine the reality of any relation. ### 8.2 Interpretation of the Continuum and Optical Emission-Line Analyses The continuum and optical emission-line analyses discussed in §3 show that SXW QSOs do differ in some respects from non-SXW QSOs. The low \[O iii\] luminosities and EWs of SXW QSOs are striking, and this phenomenon has been previously noted for a subset of SXW QSOs, the low-ionization BAL QSOs (e.g., Boroson & Meyers 1992; Turnshek et al. 1997). If \[O iii\] is an isotropic property, this result suggests that soft X-ray weakness is not merely caused by an orientation effect but arises at least in part as a result of an intrinsic property.<sup>6</sup><sup>6</sup>6Here we are following the line of reasoning discussed in §4.1 of BG92 and §3.2 of Boroson & Meyers (1992). While we recognize that \[O iii\] may have some anisotropy (e.g., see Hes, Barthel & Fobsury 1996; di Serego Alighieri et al. 1997), the best available evidence suggests that this anisotropy is not strong enough to remove the need for an intrinsic property (Kuraszkiewicz et al. 1999a). A clue to the nature of this property may be found by noting that SXW QSOs also lie toward the weak-\[O iii\] ‘negative’ end of BG92 eigenvector 1 (see §3).<sup>7</sup><sup>7</sup>7Our only SXW QSO with a ‘positive’ value of eigenvector 1 is the radio-loud QSO 1004+130, and even this object has an unusually small value of eigenvector 1 for a radio-loud QSO. A somewhat similar connection between BAL QSOs and strong optical Fe ii emitters has been discussed by Lawrence et al. (1997), although the discussion there was based on qualitative argumentation with a fairly small sample of objects. Many of the unabsorbed Seyferts and QSOs with similar values of eigenvector 1 have recently been found to have extreme X-ray spectral and variability properties, and it has been suggested that these ‘ultrasoft Narrow-Line Seyfert 1’ objects are accreting at relatively high fractions of the Eddington rate ($`\dot{M}/\dot{M}_{\mathrm{Edd}}`$; e.g., BG92; Boller, Brandt & Fink 1996; Laor et al. 1997a). If an $`\dot{M}/\dot{M}_{\mathrm{Edd}}`$ interpretation of eigenvector 1 is correct, the similar eigenvector 1 values of SXW QSOs would suggest that they also have relatively high $`\dot{M}/\dot{M}_{\mathrm{Edd}}`$. SXW QSOs and ultrasoft Narrow-Line Seyfert 1s would be related objects that differ primarily in the amount of absorption that happens to lie along the line of sight. Accretion at high $`\dot{M}/\dot{M}_{\mathrm{Edd}}`$, where radiation trapping effects are important, is expected to drive substantial mass outflow, and gas in such an outflow could produce the ultraviolet and X-ray absorption observed in SXW QSOs (see §5 of Blandford & Begelman 1999). This gas, perhaps after having cooled, may also have been detected in ultrasoft Narrow-Line Seyfert 1 objects via its emission lines; these objects appear to have especially large amounts of dense gas in their nuclei (e.g., Kuraszkiewicz et al. 1999b; Wills et al. 1999). The significant differences in the H$`\beta `$ shifts and shapes of SXW QSOs and non-SXW QSOs suggest that either (1) the dynamics of their Broad Line Regions (BLRs) are systematically different, (2) particular parts of their BLRs are preferentially viewed, or (3) broad H$`\beta `$ absorption near the line peak is modifying the profiles of SXW QSOs (e.g., Anderson 1974). While such effects might plausibly be related to the outflows discussed above, we are at present unable to find a compelling interpretation for the shift and shape differences (this is largely due to the general lack of understanding of the structure and kinematics of the BLR). The absence of obvious optical continuum reddening, while initially somewhat surprising in light of the ultraviolet and X-ray absorption discussed above, can be understood if the dust in the absorbing gas has been destroyed by sputtering or sublimation. From an empirical point of view, we also note that high-ionization BAL QSOs show only weak reddening despite their strong absorption (e.g., Weymann et al. 1991). The generally low optical continuum polarization of our SXW QSOs may also be understood by analogy with BAL QSOs. While scattering by electrons or dust, together with dust absorption of the central continuum, has been used to explain the high polarization of some BALQSOs, about half of BAL QSOs have polarization percentages $`\stackrel{<}{}1`$% (e.g., Schmidt & Hines 1999). An apparently surprising result from §7 is that $`\alpha _{\mathrm{ox}}`$ does not appear to correlate with $`R`$. Earlier studies, in contrast, have found that for a given optical luminosity, the average X-ray emission of radio-loud QSOs is $`3`$ times higher than that of radio-quiet QSOs (e.g., Z81). This apparent discrepancy is due to the following three effects: (1) Z81 define radio-loud QSOs as those having $`\mathrm{log}R>1.88`$ rather than $`\mathrm{log}R>1`$, (2) Z81 do not detect QSOs with $`\alpha _{\mathrm{ox}}<1.75`$ (see their Figure 4), and (3) the radio-loud QSOs in Z81 extend up to $`\mathrm{log}R4.8`$, while ours have $`\mathrm{log}R\stackrel{<}{}3`$. Figure 5 of Z81 suggests a flattening of $`\alpha _{\mathrm{ox}}`$ with increasing $`\mathrm{log}R`$, and to make a proper comparison with the BG92 radio-loud QSOs one needs to take subsamples from both samples which follow the same selection criteria. The average $`\alpha _{\mathrm{ox}}`$ for the subsample of 16 QSOs from Table 1 of Z81 with $`1\mathrm{log}R3`$ is $`\alpha _{\mathrm{ox}}=1.48\pm 0.05`$, which is about equal to the Z81 mean for the radio-quiet QSO population ($`1.46\pm 0.06`$) and our mean of the subsample of 14 $`\alpha _{\mathrm{ox}}>1.75`$ radio-loud QSOs from BG92 ($`1.48\pm 0.11`$). Thus, there is no discrepancy between the Z81 results and our results. The difference in $`\alpha _{\mathrm{ox}}`$ between radio-quiet and radio-loud samples appears to be driven primarily by the radio-loudest ($`\mathrm{log}R>3`$) QSOs, which are not present in our sample. ### 8.3 Future Studies Systematic hard X-ray spectroscopy of the SXW QSOs discussed in this paper is the obvious next step in the study of these objects. Such spectroscopy should allow determination of the column densities, ionization parameters, and covering factors of the expected X-ray absorbers, and it would let one critically examine the continuum of absorption properties suggested by the $`\alpha _{\mathrm{ox}}`$ versus C iv absorption EW correlation. Furthermore, improved ultraviolet spectra are needed for several of our SXW QSOs. Absorption in species in addition to C iv can constrain physical conditions in the ultraviolet absorber. Objects such as 1011–040 and 2214+139 have only weak limits on C iv absorption and might well show interesting ultraviolet absorption in higher quality spectra. Studies of these two objects will determine whether X-ray weakness in the BG92 QSOs is always associated with ultraviolet absorption, or whether it is possible in some cases to get X-ray absorption without noticeable ultraviolet absorption. Objects such as 1004+130 and 1126–041 also need better spectra to constrain the geometry and dynamics of their known ultraviolet absorption. We are working to obtain the required X-ray and ultraviolet data for our sample. Finally, the soft X-ray weakness selection method described above may be profitably applied to larger QSO samples. We thank Th. Boller, M. Elvis, E. Feigelson, J. Nousek, and D. Schneider for helpful discussions. We thank T. Boroson for providing data from BG92. We gratefully acknowledge the support of NASA LTSA grant NAG5-8107 and the Alfred P. Sloan Foundation (WNB), the fund for the promotion of research at the Technion (AL), and NASA LTSA grant NAG5-3431 (BJW).
no-problem/9908/hep-ex9908022.html
ar5iv
text
# 𝐛→𝐬⁢𝛾 Branching Fraction and CP Asymmetry ## I Introduction Because flavor changing neutral currents are forbidden in the Standard Model, electroweak penguins such as $`bs\gamma `$ give a direct look at loop and box processes. The inclusive branching fraction for $`bs\gamma `$ is important for restricting physics beyond the Standard Model, and CLEO’s 1995 published result of $`(2.32\pm 0.57\pm 0.35)\times 10^4`$ generated much theoretical interest. The branching fraction result presented here is an improved measurement with 60% additional data and enhanced analysis techniques. The theoretical branching fraction prediction from the Standard Model has also improved, with a full next-to-leading-log calculation of $`(3.28\pm 0.33)\times 10^4`$. We also introduce a new analysis: a search for CP asymmetry in $`bs\gamma `$. ## II The CLEO Detector Data for these analyses were taken with the CLEO detector at the Cornell Electron Storage Ring (CESR) with center of mass energy at the $`\mathrm{{\rm Y}}\text{(4S)}`$ resonance (10.58 GeV). Additional data were taken 60 MeV below the $`\mathrm{{\rm Y}}\text{(4S)}`$ (off-resonance) for continuum background subtraction. The CLEO detector measures charged particles over 95% of 4$`\pi `$ steradians with a system of cylindrical drift chambers. The barrel and endcap CsI calorimeters cover 98% of 4$`\pi `$, and the energy resolution for photons near 2.5 GeV in the central angular region $`(|\mathrm{cos}\theta _\gamma |<0.7)`$ is 2%. In 1995, a silicon vertex detector replaced the inner most tracking chamber, though it is not used in the analyses described here. ## III Branching Fraction Analysis This analysis is discussed in detail elsewhere and is briefly covered here. The signal for $`bs\gamma `$ is a photon from $`B`$ meson decay with $`2.1<E_\gamma <2.7`$ GeV (the published analysis used $`2.2<E_\gamma <2.7`$ GeV). The Fermi momentum of the $`b`$ quark in the $`B`$ meson and the momentum of the $`B`$ meson in the lab frame ($`B`$ mesons are produced from $`\mathrm{{\rm Y}}\text{(4S)}`$ decays with approximately 300 MeV/$`c`$ of momentum at CESR) Doppler broadens the photon line. Calculations using a spectator model indicate that 85-94% of the signal lies in the photon energy range. To obtain a photon energy spectrum, we select hadronic events with a high energy calorimeter cluster in the central region $`(|\mathrm{cos}\theta _\gamma |<0.7)`$. We reject a cluster if, when paired with another calorimeter cluster in the event, it forms a combined $`\gamma \gamma `$ mass consistent with a $`\pi ^0`$ or $`\eta `$. We also require the cluster shape be consistent with that of a single high energy photon. A naïve approach to this analysis would be to measure the photon spectrum in on-resonance data and use the off-resonance data to subtract the continuum background. But the backgrounds from continuum with initial state radiation ($`e^+e^{}q\overline{q}\gamma `$) (ISR) and from continuum processes ($`e^+e^{}q\overline{q}`$) involving a high energy $`\pi ^0`$, $`\eta `$, or $`\omega `$ where one of the daughter photons is not detected dominate to such an extent that extraction of a signal is impossible. We therefore suppress the continuum background with two separate methods and subtract what remains with off-resonance data. The first continuum suppression scheme involves exploiting differences in the event shapes between signal ($`B\overline{B}`$) and continuum. While continuum events appear jetty and continuum with ISR appear jetty when transformed to the rest frame of the system after the initial state radiation, the signal events are spherical due to the other $`B`$ meson decaying in the event. We characterize the event shape with eight variables: the Fox-Wolfram second moment ($`R_2`$), ratio of the sum of transverse momenta of particles more than 45 away from the high energy photon axis to the total scalar sum of particle momenta in the event excluding the momentum of the high energy photon ($`S_{}`$), $`R_2`$ transformed to the rest frame of the system after ISR (assuming the high energy photon in the event is initial state radiation) ($`R_2^{}`$), $`\mathrm{cos}\theta `$ between the photon and the thrust axis of the rest of the event transformed into the rest frame of the system after ISR ($`\mathrm{cos}\theta ^{}`$), and the energies of particles found in 20 and 30 cones parallel and anti-parallel to the high energy photon direction. Since no one variable provides enough discriminating power, we combine them with a neural network into a single variable $`r`$. The distribution of $`r`$ tends towards $`1`$ for continuum and ISR events and $`+1`$ for signal $`bs\gamma `$ events. This shape method is approximately 30% efficient for signal. The other method of suppressing continuum and ISR events is a pseudo-reconstruction of the parent $`B`$ meson involved in the $`bs\gamma `$ decay. We look for a charged or neutral kaon with up to four pions, one of which may be a $`\pi ^0`$, which when combined with the high energy photon are consistent with being the decay products of a $`B`$ meson. We try all combinations of kaons and pions given the above restrictions and choose the one that minimizes a $`\chi ^2`$ that includes, $$\chi _B^2=\left(\frac{MM_B}{\sigma _M}\right)^2+\left(\frac{EE_{\text{beam}}}{\sigma _E}\right)^2$$ (1) as well as contributions from particle identification such as $`dE/dx`$ and $`\pi ^0`$ and $`K_s^0`$ mass deviations. In Eq. (1), $`M=\sqrt{E_{\text{beam}}^2P^2}`$ is the beam-constrained mass, where $`P`$ is the vector sum of particle momenta comprising the $`B`$ meson candidate, $`M_B`$ is the $`B`$ mass (5.28 GeV/c<sup>2</sup>), $`E`$ is the energy of the candidate $`B`$, and $`\sigma _M`$ and $`\sigma _E`$ are the resolutions on the beam constrained mass and the measured $`B`$ energy respectively. Events that have a minimum $`\chi _B^2<20`$ are considered pseudo-reconstructed (“pseudo” is used because this method is only for continuum suppression, and we are not concerned that the parent $`B`$ reconstruction be exactly correct). If the minimum $`\chi _B^2>20`$ or not one combination of kaons and pions was found in the event, then the shape analysis described above is used. For events that are pseudo-reconstructed, we calculate the angle between the thrust axis of the inferred signal B meson and the thrust axis of the rest of the event $`(\mathrm{cos}\theta _{tt})`$. Events that are in actuality continuum typically have $`|\mathrm{cos}\theta _{tt}|`$ peaked at 1, while signal events have a flat distribution. To get the most discriminating power, we combine $`\chi _B^2`$ and $`\mathrm{cos}\theta _{tt}`$ from the pseudo-reconstruction and $`r`$ from the shape neural network into a new variable $`r_c`$ using another neural network. As with $`r`$, the distribution of $`r_c`$ tends towards $`+1`$ for signal events and $`1`$ for continuum. While this combined method is 10% efficient for signal, it reduces the continuum background by an additional factor of four over the shape method alone. We then weight each event according to its value of $`r_c`$ if it is pseudo-reconstructed or $`r`$ if not. The weighting scheme is designed so that the weighted yield is equivalent to the event yield in the absence of continuum backgrounds. This procedure gives the smallest statistical uncertainty on this background subtracted yield. Note that there are non-continuum backgrounds discussed below, but these are small compared to continuum. The continuum background remaining after the suppression efforts is determined from off-resonance data. $`B\overline{B}`$ backgrounds are dominated by $`bc`$ (e.g. $`BX\pi ^0`$ and $`BX\eta `$) though $`bu`$ and $`bsg`$ may have a small contribution. We estimate the $`B\overline{B}`$ backgrounds with Monte Carlo (MC), but correct the $`\pi ^0`$ and $`\eta `$ momentum spectra for any differences between data and MC. We obtain the momenta spectra from data by treating $`\pi ^0`$’s and $`\eta `$’s as if they were photons and follow the same analysis procedure using weights described above. Thus, the MC is used only for the $`\pi ^0`$ and $`\eta `$ veto efficiencies and for any other small $`B\overline{B}`$ backgrounds not originating from missing a $`\gamma `$ daughter from a $`\pi ^0`$ or $`\eta `$. The weighted yields from 3.1 fb<sup>-1</sup> of on-resonance data plotted with respect to photon energy are shown in Fig. 1, with all of the backgrounds shown in (a) and the background subtracted weighted yield, which is the signal, shown in (b). 1.6 fb<sup>-1</sup> of off-resonance data were used for the continuum subtraction. The expected spectrum for $`bs\gamma `$ from a spectator model is also shown in (b). In the region of interest ($`2.1<E_\gamma <2.7`$ GeV) we measure $`500.5\pm 7.4`$ on-resonance weighted events, $`382.3\pm 7.0`$ scaled off-resonance weighted events, and estimate the $`B\overline{B}`$ background to be 20.6 weighted events (all uncertainties statistical). Applying the $`\pi ^0`$ and $`\eta `$ corrections to the estimate ($`1.9\pm 0.5`$ and $`3.6\pm 1.1`$ weighted events respectively), we derive the weighted, background-subtracted yield of $`92.2\pm 10.3\pm 6.5`$, where the first uncertainty is statistical and the second is systematic. For a typical sample of MC $`bs\gamma `$ events, approximately 50% pass all analysis requirements and of these, 40% are pseudo-reconstructed ($`\chi _B^2<20`$). But since we use weighted events for the signal yield, we need a weighted efficiency; the sum of weights passing all requirements. Because of our definition of weights, the weighted efficiency tends to be smaller than the efficiencies stated above. To model the signal, we use the spectator model of Ali and Greub, which includes gluon bremsstrahlung and other higher order radiative effects. Details of MC production and parameters used are given in ref . The weighed efficiency is $`(4.43\pm 0.29\pm 0.22\pm 0.03\pm 0.31)\times 10^2`$, where the first uncertainty is due to spectator model inputs, the second due to recoil system hadronization, the third due to uncertainty in the production ratio of neutral and charged $`B`$ meson pairs, and the last to detector modeling. The weighted efficiency is combined with the background subtracted weighted yield to give the inclusive branching fraction of $`bs\gamma `$ as $`(3.15\pm 0.35\pm 0.32\pm 0.26)\times 10^4`$, where the uncertainties are statistical, systematic and for model dependence respectively. Our results are in agreement with the Standard Model prediction. Conservatively allowing for the systematic uncertainty, we find that the branching fraction must be between $`2.0\times 10^4`$ and $`4.5\times 10^4`$ (each limit at 95% CL). ## IV CP Asymmetry The SM predicts no CP asymmetry in $`bs\gamma `$ decays, but some recent theoretical work suggests that non-SM physics may significantly contribute to a CP asymmetry. Furthermore, if new physics has a weak phase difference near 90 with respect to the SM and the strong phase is non-zero, the new physics would only slightly alter the $`bs\gamma `$ inclusive branching ratio, but could produce a large effect in the asymmetry. Since we can reuse much of the machinery from the branching fraction measurement, a search for CP asymmetry is an obvious extension to the $`bs\gamma `$ analysis. The pseudo-reconstruction procedure described above is used to tag the $`b`$ quark flavor ($`b`$ or $`\overline{b}`$), because it determines the particles decayed from the signal $`B`$ meson. But since the pseudo-reconstruction was designed for continuum suppression, we must determine how often it reconstructs the $`b`$ flavor correctly. If a signal $`B`$ meson decays to a neutral kaon and overall neutral pions (e.g. $`B^0K_s^0\pi ^+\pi ^{}`$), then we cannot determine the $`b`$ flavor from pseudo-reconstruction. But for the majority of cases where the $`b`$ flavor is determinable, it can be deduced by the charge of the kaon, or the sum pion charge if the kaon is neutral. From MC, we estimate the probability of reconstructing the incorrect $`b`$ flavor when the $`b`$ flavor is determinable to be $`8.3\pm 1.6\%`$. The quoted uncertainty includes a combination of MC statistics and uncertainties derived from varying spectator model inputs to the MC and varying the ratio of charged to neutral $`B`$ meson decays. We also take into account the probability of tagging the $`b`$ flavor when it should be indeterminable and vice–versa, though these effects are small and will not be discussed further here. The asymmetry we measure is $`(N_1N_2)/(N_1+N_2)`$, where $`N_1`$ is the weighted yield tagged as $`b`$ quarks from pseudo-reconstruction and determinable as such, and $`N_2`$ is similarly the weighted yield tagged as $`\overline{b}`$. From the misreconstruction rates described above, we apply a multiplicative correction factor of $`1.22\pm 0.04`$ to the asymmetry. In our misreconstruction determination, we assumed that the rate of mistagging a $`b`$ as a $`\overline{b}`$ was the same as mistagging a $`\overline{b}`$ as a $`b`$. Initial studies lead us to apply a conservative additive systematic error on the asymmetry of 5% to account for unequal mistagging rates, though we observe no measurable asymmetry in the off-resonance data and other control samples. Using the weighted yield from pseudo-reconstructed events with photons within $`(2.2<E_\gamma <2.7\text{ GeV})`$, we observe $`36.8\pm 5.2`$ weights reconstructed as $`b`$ quarks and $`28.4\pm 5.2`$ weights reconstructed as $`\overline{b}`$ quarks (uncertainties are statistical only), leading to a raw asymmetry measurement of $`0.13\pm 0.11`$. Applying the correction factors, we obtain our preliminary CP asymmetry measurement in $`bs\gamma `$ decays: $`(0.16\pm 0.14\pm 0.05)\times (1.0\pm 0.04)`$, where the first number is the central value, the second is the statistical uncertainty, the third is the additive systematic described above, and the multiplicative factor is from the uncertainty on the mistagging rate correction. Within the uncertainties, we measure no CP asymmetry. From this result, we derive 90% CL limits on the CP asymmetry ($`𝒜`$) of $`0.09<`$ $`𝒜`$ $`<0.42`$. ## V Summary and Outlook We presented CLEO’s preliminary results on the $`bs\gamma `$ branching ratio and the CP asymmetry in $`bs\gamma `$ decays with approximately 3.1 fb<sup>-1</sup> of on-resonance data. CLEO has since accumulated a total of 10 fb<sup>-1</sup> of on-resonance data. Work is underway to determine the $`bs\gamma `$ inclusive branching fraction and CP asymmetry with all of the data. We are also exploring extending both $`bs\gamma `$ analyses by tagging the $`b`$ flavor with a lepton from the other $`B`$ meson in the event it undergoes semileptonic decay. This procedure offers further continuum suppression and may improve the asymmetry measurement by allowing the use of events where the $`b`$ flavor is undeterminable by pseudo-reconstruction. We gratefully acknowledge the effort of the CESR staff in providing us with excellent luminosity and running conditions. We also thank M. Neubert for bringing the theoretical interest in $`bs\gamma `$ CP asymmetry to our attention. J.R. Patterson and I.P.J. Shipsey thank the NYI program of the NSF, M. Selen thanks the PFF program of the NSF, M. Selen and H. Yamamoto thank the OJI program of the DOE, J.R. Patterson, K. Honscheid, M. Selen and V. Sharma thank the A.P. Sloan Foundation, M. Selen and V. Sharma thank the Research Corporation, F. Blanc thanks the Swiss National Science Foundation, and H. Schwarthoff and E. von Toerne thank the Alexander von Humboldt Stiftung for support. This work was supported by the National Science Foundation, the U.S. Department of Energy, and the Natural Sciences and Engineering Research Council of Canada.
no-problem/9908/astro-ph9908165.html
ar5iv
text
# Early results of the ESO VLT ## 1. Introduction The general goals that ESO intended to reach with Science Verification (SV) Observations were manifold (Giacconi et al. 1999): i) to experiment with scientific observing runs, including the end-to-end VLT dataflow (Silva and Quinn, 1997), ii) to produce data of scientific quality, involving as early as possible the ESO community in their analysis, iii) to foster an early scientific return from the VLT, iv) to obtain feedback about the telescope performances and operational procedures. To this end an SV team has been formed at ESO, under the leadership of A.Renzini. A first block of SV observations has been carried out in August 1998 with the VLT test camera. Results have been published in Vol. 343 of the A&A Journal. A second block of SV observations was planned to be executed in January and February 1999 and in particular with the two instruments for optical/IR imaging and spectroscopy: FORS1 (Nicklas et al. 1997) and ISAAC (Moorwood 1997). The FORS-1/ISAAC SV Team includes the following scientists: J. Alves, S. Cristiani, R. Hook, R. Ibata, M. Kissler-Patig, P. Møller, M. Nonino, B. Pirenne, R. Rengelink, A, Renzini, P. Rosati, D. Silva, E. Tolstoy, and A. Wicenec. The actual schedule of the SV observations was more complex than originally planned and benefitted from observations executed by the FORS1 and ISAAC Commissioning Teams. FORS1 and ISAAC are rather complex instruments, with many observing modes. The goal of the SV has been to cover the main modes, selecting scientific programmes of outstanding interest. The following observations are briefly described here: 1. The Cluster Deep Field MS1008.1-1224 2. The Antlia dwarf spheroidal Galaxy 3. Multiple Object Spectroscopy (MOS) of Lyman break galaxies in the AXAF and Hubble Deep Field South 4. ISAAC IR Spectroscopy of the gravitationally magnified galaxy at z=2.72 MS1512cB58. ## 2. The FORS-ISAAC Cluster Deep Field MS1008.1-1224 The criteria inspiring the selection of the cluster were: existence of published data indicating a large mass/velocity dispersion at redshift=$`0.30.5`$, existence of gravitational arcs, optimal visibility throughout most of the night in January-March (i.e. RA=8-11hr). The selection narrowed on the X-ray selected cluster MS1008.1-1224 (RA = 10 10 32.2, DEC =–12 39 55. ep.2000) from the Einstein Medium Sensitivity Survey (EMSS, Gioia and Luppino, 1994) at z=0.30, also part of the CNOC Survey (Carlberg et al. 1996). BVRIJK observations were carried out with total integration times between 4000 and 5400 s in the optical and about 1h in the IR. The K-band observations were obtained under exceptional seeing conditions and produced de-jittered coadded images of $`0.4^{\prime \prime }`$ PSF. These data will be used for: a) a detailed study of the cluster mass distribution from gravitational lensing shear maps, magnification bias and strong lensing features, b) a study of the cluster galaxy population down to $`4`$ mag below L, using color-mag diagrams, c) a search for highly magnified distant galaxies, d) obtaining photometric redshifts of all galaxies in the field, e) identifying interesting stars in the field. ## 3. The Antlia Dwarf Spheroidal Galaxy Antlia is a small low surface brightness dwarf spheroidal type galaxy (RA= 10 04 04, DEC=–27 19 49, ep. 2000, l=263, b=22) first noted in an HI survey of southern hemisphere galaxies by Fouqué et al. (1990), which was discovered in 1997 to be an outlying member of the Local Group. This galaxy was selected for the SV because: published data indicate a distance close enough to allow the detection of resolved stars well down the Red Giant Branch, there is a controversy about the existence and importance of a young stellar population and reddening in Antlia and the visibility is optimal in January-March. Images have been obtained with FORS in B ($`4\times 600s`$), V ($`4\times 600s`$) and I ($`18\times 300s`$), all with a PSF close to $`0.5^{\prime \prime }`$. The combined color image is shown in Fig. 2. These data will be used for: i) a detailed study of the Colour-Magnitude diagrams to determine the properties of the resolved stellar population in this small nearby, relatively isolated galaxy, ii) determining if there is a young population from the presence or absence of a Main Sequence, iii) comparing the properties of a more distant dwarf spheroidal galaxy with those we see around our Galaxy, iv) a study of the reddening properties of this system with colour-colour diagrams, v) confirming the distance to this system. ## 4. Multiple Object Spectroscopy of Lyman break galaxies in the AXAF and Hubble Deep Field South The study of Lyman-break galaxies was the main spectroscopic programme of the FORS1 Science Verification. The science goals are: i) to push the FORS1 MOS to the limits and optimize its observational strategy and data reduction procedures for the specific case of faint galaxies, ii) to provide pilot observations for future studies aiming at characterizing the abundance of high-redshift galaxies by luminosity, size, morphology, star formation rate and clustering, iii) to check and refine the photometric selection criteria for high-z galaxies, iv) help to tune future wide-angle optical surveys to maximize the yield of Lyman-break galaxies, v) to provide galaxies in suitable redshift ranges for future spectroscopic observations with ISAAC. The AXAF field (Giacconi et al., 1998) was selected for SV observations, given the availability of deep multi-color imaging obtained with SUSI-2 and SOFI and derived catalogues produced by the EIS project. Data of the same type were also publicly available for a field including the WFPC-2 pointing in the Hubble Deep Field South. Targets in this latter field were provided to the FORS1 commissioning team and were observed by them in December 1998. Targets were extracted from the lists of U and B dropouts given in Table 5 and Table 6 of Rengelink, R. et al. (1998) for the AXAF field and in Table 9 of da Costa, L. et al. (1998) for the HDF-S. When no suitable candidate was available for the allowed range of positions of a given slit, a random object in the field was chosen. Spectra of the objects in both the AXAF field and HDF-S have been reduced and analyzed in the same way within the MIDAS package. A total of 8 galaxies with redshifts between 2.8 and 4 were identified. The spectra of two of them are shown in Fig. 3. A full table and details of the reduction are given at the URL http://http.hq.eso.org/science/ut1sv/MOS\_index.html. ## 5. Low-resolution spectroscopy in the H band of MS1512-cB58 The gravitationally lensed galaxy MS1512-cB58 was serendipitously discovered by Yee et al. (1996), in the course of the CNOC cluster survey. It lies in the field of the z=0.36 cluster MS1512+36, and its redshift was derived from a dozen strong absorption lines in the rest frame UV (Ellingson et al. 1996). Most interestingly, this galaxy is gravitationally amplified by a factor of about 30. Its near-IR magnitudes are J=19.12, H=18.42, K’=17.83, i.e. the object has exceptionally bright apparent magnitudes for its redshift. Therefore, it offers a unique opportunity to get a high S/N infrared spectrum of a high redshift galaxy, with important lines such as $`H_\alpha `$ (at the edge of the K band) and $`H_\gamma `$ (within the H band) being uncontaminated by atmospheric emission/absorption. The detection of such lines would allow a more reliable determination of the star formation rates, compared to other indicators such as \[O II\] and the UV continuum. More importantly, the profile of such lines - if present - may provide hints on the mass of this high-redshift galaxy. ISAAC is the first instrument that offers this opportunity, and the unusually large amplification factor made it an attractive target for the VLT, in spite of the large zenithal distance at which it can be observed from Paranal. The data were obtained in the night of March 24, 1999. ISAAC was operated in the short-wavelength, low-resolution mode. The SH filter was used (covering 1.4 to 1.82 micron), with the $`1^{\prime \prime }`$ slit. The resolution in this configuration is around 500. The pixel scale is $`0.147^{\prime \prime }`$/pixel. A preliminary data reduction was carried out with a final exposure of 5040 sec effective exposure time. The result is shown in Fig. 4. ## 6. Data Access The data have been made public for the ESO community and, in the case of HDF-S, worldwide. Full details and data request forms can be found at the URL: http://http.hq.eso.org/science/ut1sv/. ## References Carlberg, R.G., Yee, H.K.C., Ellingson, E.; Abraham, R., Gravel, P., Morris, S., Pritchet, C. J., 1996, ApJ, 462, 32 da Costa, L. et al. 1998, A&A submitted, astro-ph/9812105 Ellingson, E., Yee, H.K.C., Bechtold, J., Elston, R., 1996, ApJ 466, L71 Fouqué, P., Durand, N., Bottinelli, L., Gouguenheim, L., Paturel, G., 1990, A&AS 86, 473 Giacconi R et al., Proceedings of ”Highlights of X-ray Astronomy”, in honour of J.Truemper’s 65th birthday Garching, Germany, 17-19 June 1998 Ed. B.Aschenbach Giacconi R., Gilmozzi R., Leibundgut B., Renzini A., Spyromilio J., Tarenghi M., 1999, A&A 343, L1 Gioia I., Luppino, G.A., 1994, ApJS 94, 583 Moorwood A., 1997, SPIE 2871, 1146 Nicklas, H., Seifert, W., Boehnhardt, H., Kiesewetter-Koebinger, S., Rupprecht, G., 1997, SPIE 2871, 1222 Rengelink, R. et al. 1998, A&A submitted, astro-ph/9812190 Silva D., Quinn, P., 1997, The Messenger 90, 12 Yee, H.K.C., Ellingson, E., Bechtold, J., Carlberg, R.G., Cuillandre, J.-C., 1996, AJ, 111, 1783
no-problem/9908/astro-ph9908341.html
ar5iv
text
# Far Ultraviolet Absolute Flux of 𝛼 Virginis1footnote 11footnote 1Based on the development and utilization of the Espectrógrafo Ultravioleta de Radiación Difusa, a collaboration of the Spanish Instituto Nacional de Técnica Aeroespacial and the Center for EUV Astrophysics, University of California, Berkeley ## 1 Introduction Observations of OB stars in the far-ultraviolet (FUV), below 1100 Å, are important to address several issues, such as the interaction of these stars with their surrounding ISM, or the determination of interstellar dust size. These observations can also serve as a test of stellar atmosphere models and to set spectrophotometric standards in the FUV, which are brighter and more abundant than white dwarfs, for future space observatories. There are very few FUV spectroscopic observations of stars. Since the first rocket observations of stellar emission in the FUV (Brune, Mount, & Feldman 1979), several attempts have been made to determine the spectral energy distribution of stars of different spectral types in this wavelength range. In early work rockets were used to study stars in the FUV, and there were some spacecraft observations by Voyager (Broadfoot et al. 1977) and Copernicus (Rogerson et al. 1973) spacecraft observations. More recently, three instruments with FUV spectrometers, HUT (Davidsen 1990), ORFEUS (Hurwitz & Bowyer 1991) and UVSTAR (Stalio et al. 1993), have flown on-board the Space Shuttle. Absolute calibration in this wavelength range is difficult due to the absence of primary or secondary standards, and indirect calibrations were applied in all cases. The consequence is that large flux differences, of up to a factor of ten at certain wavelengths, are found between observations of the same star taken with different instruments, and none of them agrees with Kurucz model atmospheres below 1000 Å. Voyager (Chávez, Stalio, & Holberg 1995) and HUT (Buss, Kruk, & Ferguson 1995) observations report fluxes higher than the model predictions below 1200 Å. On the other hand, rocket observations (Brune et al. 1979; Carruthers, Heckathorn, & Opal 1981; Woods, Feldman, & Bruner 1985; Cook, Cash, & Snow 1989) gave substantially lower fluxes than those from Voyager, and also lower than the fluxes predicted by Kurucz model atmospheres. $`\alpha `$ Vir is one of the most studied stars in the FUV. It has been observed by Copernicus (York & Kinahan 1979), Voyager 1 and 2 (Holberg et al. 1982), and by rockets (Brune et al. 1979; Cook et al. 1989, and Wilkinson et al. 1995). The fundamental parameters of $`\alpha `$ Vir are well known. Spica is a double-lined spectroscopic binary with an ellipsoidal variation of 0.03 mag. due to tidal distortion with 4-day orbital period, superposed to $`\beta `$ Cephei-type pulsations of the primary, of 0.016 mag. amplitude and a period of $`0\stackrel{\mathrm{d}}{\mathrm{.}}1738`$ (Shobbrook et al. 1969). The radius of the primary, determined interferometrically by Herbison-Evans et al. (1971), is $`R_1=8.1\pm 0.5`$ $`R_{}`$ for a distance of $`84\pm 4`$ pc, and its spectral type is B1V. The secondary component of $`\alpha `$ Vir is probably a B4V star, for which Popper (1980) assumed a $`(BV)=0.18`$, leading to a radius of $`R_2=4.16\pm 1.17`$ $`R_{}`$ which is consistent with the mass ($`10.9\pm 0.9`$ $`M_{}`$) deduced by Herbison-Evans et al. (1971). Its parallax has been measured by Hipparcos (12.44$`\pm `$ 0.86 mas, Perryman et al. 1997). Corrections for the radii of the component stars of $`\alpha `$ Vir due to this new value of the distance lead to R<sub>1</sub>=7.78 and R<sub>2</sub>=3.99. Integrated photometry for the two components from the literature as extracted from the SIMBAD database indicate $`V=0.98`$, $`BV=0.235`$, and $`UB=0.94`$ for Johnson photometry and $`by=0.114`$, m<sub>1</sub> = 0.080, and c<sub>1</sub> = 0.018 for Stromgren uvby photometry. Crawford H$`\beta `$ photometry gives $`\beta `$ = 2.607. In this paper we present the spectrum of $`\alpha `$ Vir obtained with EURD (Espectrógrafo Ultravioleta extremo para la Radiación Difusa) on-board MINISAT-01. We compare our observations with previous ones and with Kurucz models (ATLAS9, Kurucz 1993). Our observations allowed us to obtain a flux calibrated spectrum of $`\alpha `$ Vir with the best signal-to-noise ratio and spectral resolution to date in the FUV range. With this observations we intend to determine the absolute flux of $`\alpha `$ Vir in the FUV, for which discrepancies have been found in previous works. ## 2 Observations and data reduction EURD was launched on April 1997 on-board the Spanish satellite MINISAT-01, which has a retrograde orbit of 151 inclination and an altitude of 600 km. Details on the mission can be found in Morales et al. (1998). EURD is a spectrograph specially designed to observe diffuse radiation in the wavelength range from 350 to 1100 Å. It observes in the anti-sun direction and during orbital eclipse. A precise description of the instrument and its ground calibration can be found in Bowyer, Edelstein, & Lampton (1997). The detector is a photon counter device that produces spectral images, with spatial resolution capabilities along an axis perpendicular to the direction of spectral dispersion. The spectral resolution of the instrument is $`5`$ Å. When a bright, early-type star falls within the $`25^{}\times 8^{}`$ field of view, it shows up as emission longward of $`912`$ Å, on a finite area along the spatial dimension. Given the mission pointing constraints, only stars within $`13^{}`$ from the ecliptic can be observed. Observations of $`\alpha `$ Vir presented in this paper were taken from 1998 April 13 to April 19, and 1999 April 2 to April 23. During the data reduction process, we tracked the position of the stellar emission on the detector for every second, extracted the photons detected within 10$`\stackrel{}{\mathrm{.}}`$5 from the emission maximum (to include all the stellar emission gathered during one second), and subtract a background from an area of the detector close to where the stellar emission lies. Photon counts are then corrected by the efficiency of the detector as a function of incidence angle. The accumulated photon counts are converted to fluxes by applying the in-flight calibration performed with simultaneous observations of the full Moon with EUVE and EURD (Edelstein et al., in preparation). ## 3 Results and discussion ### 3.1 The Far-UV spectrum of $`\alpha `$ Vir Fig. 1 shows the spectrum of $`\alpha `$ Vir , with a total integration time of $`1.06\times 10^5`$ s, and covering the wavelengths $`\lambda <1080`$ Å. The noise level of the spectrum is $`\sigma 3.3\times 10^{11}`$ erg<sup>-1</sup> cm<sup>-2</sup> Å<sup>-1</sup> and the signal to noise ratio is $`>3000`$. Some absorption features are obvious in the spectrum: the Lyman series of hydrogen at 937, 949, 972 and 1026 Å, N III at 989 Å, S III at 1012 Å, C II and O IV blended at 1037 Å, and an unidentified feature at $``$1063 Å (see York & Kinahan 1979). ### 3.2 Comparison with model atmospheres Atlas 9 Kurucz models properly reproduce the UV and optical spectra of B stars (Malagnini et al. 1985; Fitzpatrick & Massa 1998; Chávez et al. 1995). For wavelengths below 1200 Å, Holberg et al. (1982) reported for $`\alpha `$ Vir a flux excess with respect to Kurucz models which is within the range of uncertainty of the reddening. They used a model atmosphere combination of 24500 K and 17000 K, and a reddening of $`E(BV)=0.02`$. In this wavelength range Buss et al. (1995) also found a flux excess of $`5\%`$ at 1000 Å with respect to Kurucz models in their sample of galactic OB stars. In this work, we first compared Kurucz models with IUE spectra of $`\alpha `$ Vir and then we checked if the model extension to our wavelengths properly fits the EURD spectra. The FUV flux recorded by EURD is the combined flux of the two components of $`\alpha `$ Vir binary system. Therefore it is necessary to build a combined Kurucz model corresponding to the $`\alpha `$ Vir system. For this purpose we have adopted the effective temperatures given by Popper (1980), T<sub>1</sub> = 24500 and T<sub>2</sub> = 17200. We computed the model used for the fit as a combination of Kurucz models of these temperatures (taking into account the radii of the two components), and gravity $`\mathrm{log}g=3.69`$. Spica IUE spectra (SWP33091HL and LWR13650HL) have been selected from the INES database. To create a single spectrum, we used the SWP spectrum shortward of 1940 Å, and the LWP one for longer wavelengths. The IUE data were degraded to the 10 Å spectral resolution of the models. Correction for interstellar HI and H<sub>2</sub> absorption are negligible, given the low hydrogen column density (Fruscione et al. 1994) in front of this relatively close star. We normalized the resultant Kurucz model to the IUE spectrum. Adjusting to IUE observations we avoid the uncertainties possibly present in the determination of the angular radius. Fig. 2 shows the Kurucz model scaled to IUE, together with the EURD spectrum applying no reddening correction, binned down to match the spectral resolution of the models. The flux of the observed spectrum is $`20`$% higher than the model flux. Fig. 3 shows the comparison of the Kurucz model with both EURD and IUE spectra. We have also checked how a possible color excess could affect our results. The reddening correction is crucial at this wavelength range. Even though $`\alpha `$ Vir has a very low optical reddening, at FUV wavelengths the extinction rises very steeply and must be corrected very carefully for an absolute flux determination. For wavelengths longer than 1200 Å the interstellar extinction has been well studied (Savage & Mathis 1979; Seaton 1979; Cardelli, Clayton and Mathis 1989; Fitzpatrick & Massa 1990). However, for wavelengths below 1200 Å, few attempts have been done to determine an extinction law (Longo et al. 1989; Snow, Allen, & Polidan 1990; Buss et al. 1994). The extinction laws obtained agree in shape with the extrapolation of Cardelli et al. (1989) law, but their absolute values are very dependent of the value of the ratio of total to selective extinction, $`R_v`$, in the star direction. We have used the extrapolation of the average extinction law of Cardelli et al. (1989), applying $`R_v=3.1`$ for the diffuse interstellar medium, which is the average of the two more recent and accurate determinations of mean values of $`R_v`$ (3.08, He et al. 1995; 3.12, Whittet & van Breda 1980). We have followed three different methods to obtain $`E(BV)`$: column density of hydrogen, Strömgren uvby photometry, and Johnson UBV photometry. The highest value of color excess for $`\alpha `$ Vir ($`E(BV)=0.02`$) was obtained using Johnson UBV photometry and the intrinsic colors of Schmidt-Kaler (1982), taking the observed $`(\mathrm{B V})=0.235\pm 0.08`$ as derived from the recent ground UBV photometry carried out for the Hipparcos Mission (Perryman et al. 1997). This color excess of 0.02 is the same used by Holberg et al. (1982). Using this value to deredden IUE and EURD spectra, and scaling the corresponding Kurucz model to the new IUE flux, we find that EURD fluxes in this case would be in excess by 30% with respect to the model expectations. Since the flux excess using $`E(BV)=0`$ was 20%, we can see that, in this case, the redenning correction is not critical in order to compare our results with the models. We conclude that the data are reasonably consistent with Kurucz models both in flux and spectral shape. A fit to within $`5\%`$, would be obtained by increasing the adopted temperature of the primary by 600 K and assuming no reddening. ### 3.3 Comparison with previous observations There have been very few observations of stellar spectra in the FUV. Among them, $`\alpha `$ Vir is one of the most studied stars in this range (Brune et al. 1979; Holberg et al. 1982; Cook et al. 1989; Wilkinson et al. 1995). An indication of the difficulties involved in this study is that there exist significant differences in the flux derived for this star by different authors. Cook et al. (1989) and Brune et al. (1979) spectra are mutually consistent, but they are lower than those taken by Voyager (Holberg et al. 1982) and recently by Wilkinson et al. (1995) with a sounding rocket. Fig. 4 shows the EURD spectrum compared with previous observations of $`\alpha `$ Vir. The fluxes we derive are similar to those obtained by Holberg et al. (1982). By degrading our spectral resolution to match that of Holberg et al. (1982) observations, we see only a small discrepancy in the Lyman lines of hydrogen, whose absorption is deeper in the EURD spectrum. This can be due to residual atmospheric extinction in our data, which is not present in Voyager observations. Note, however, that the depth of the Lyman absorption lines in EURD data are more consistent with the prediction of the Kurucz models (Fig. 2). As for the flux discrepancy between previous observations, we suggest that the fluxes obtained by EURD and Voyager better represent the real far-UV flux of $`\alpha `$ Vir, while the group of observations that provide lower fluxes (Cook et al. 1989; Brune et al. 1979) seems to underestimate it. This suggestion is supported by the better consistency between our data and those of IUE, since the EURD spectrum is close to the Kurucz model that better reproduce IUE data (Fig. 2). The situation could switch in favor of the lower-flux observations if Kurucz models proved to overestimate the fluxes in this wavelength region. ## 4 Conclusions We present new $`\alpha `$ Vir observations below 1080 Å with improved spectral resolution and signal to noise ratio, taken with the EURD spectrograph on-board MINISAT 01. We compared the EURD $`\alpha `$ Vir spectrum with a Kurucz model atmosphere computed with the best values of the temperature, distance and radius of the components of $`\alpha `$ Vir binary system. This comparison shows that models are in reasonable agreement with the flux measured by EURD, being 20-30% higher than the models. Our results support Voyager fluxes of Holberg et al. (1982), rather than the lower fluxes given by rocket observations. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France, and of the International Ultraviolet Explorer data retrieved from the INES Archive. The development of this instrument has been partially supported by INTA grant IGE 490056. C.M. and J.T. acknowledge support by DGCYT grant PB94-0007. JFG is supported in part by DGICYT grant PB95-0066 and by Junta de Andalucía (Spain). Support for the publication of this paper has been provided by INTA. The UCB authors wish to thank Yumi Odama for help in the data processing. Partial support for the development of the EURD instrument was provided by NASA grant NGR 05-003-450. When NASA funds were withdrawn by Ed Weiler, the instrument was completed with funds provided by S. Bowyer. The UCB analysis and interpretation work is carried out through the volunteer efforts of the authors.
no-problem/9908/quant-ph9908087.html
ar5iv
text
# Untitled Document Quantum particle constrained to a curved surface in the presence of a vector potential. Mario Encinosa and Ray H. O’Neal Department of Physics Florida A$`\&`$M University, Tallahassee, Florida Abstract The Schrodinger equation for a charged particle constrained to a curved surface in the presence of a vector potential is derived using the method of forms. In the limit that the particle is brought infinitesimally close to the surface, a term arises that couples the component of the vector potential normal to the surface to the mean curvature of the surface. PACS numbers: 03.65.Ge,3.65 -w It is often stated that the quasifree electrons of a nanostructure are confined to very nearly two dimensional regions . A heterojunction well, for one example, confines electrons to a layer much smaller than the lateral dimensions of the device , so that the physics of the object may be discussed in terms of standard two dimensional quantum mechanical models . Here we are interested if the reduction to lower dimensionality (even if only approximately realized) via constraints can result in novel effects when a charged particle is subject to a static vector potential. There are none for a flat surface from elementary considerations; however, it has been shown that when a particle in three dimensional space is constrained to a two dimensional curved surface, curvature terms enter the Schrodinger equation (or the path integral) for the particle . In this letter, the methods of are extended to include a vector potential. Let $`d`$ be the standard exterior derivative operator that provides a mapping from $`p`$ forms to $`p+1`$ forms appropriate to a coordinate system $`(q_1,q_2,q)`$. The minimal substitution rule for the covariant derivative of a negatively charged particle in the presence of a vector potential gives $$\frac{1}{i}D\frac{1}{i}\left[d+ieA\right]$$ $`(1)`$ In the absence of a scalar potential, A is $$A=A_1\sigma _1+A_2\sigma _2+A_3\sigma _3$$ $`(2)`$ with $`\sigma _i`$ one forms derived from applying $`d`$ to the Monge form $`𝐱(q_1,q_2,q)`$ for a given coordinate system. For $`\mathrm{\Psi }`$ a zero form function, the Schrodinger equation in this notation (in natural units) becomes $$\frac{1}{2}[DD\mathrm{\Psi }]=i\mathrm{}\frac{\mathrm{\Psi }}{t}$$ $`(3)`$ with $``$ the Hodge star operator. In a gauge with $`dA=0`$, eq.(3) may be written in the familiar form $$\frac{1}{2}\left[dd+2ieAd(AA)\right]\mathrm{\Psi }=i\mathrm{}\frac{\mathrm{\Psi }}{t}.$$ $`(4)`$ Because the main interest here is in regions near to and on surfaces, $`𝐱`$ is set equal to a surface part with coordinates $`(q_1,q_2)`$ plus a part everywhere normal to the surface with coordinate $`q`$, $$𝐱(q_1,q_2,q)=𝐫(q_1,q_2)+q𝐞_3$$ $`(5)`$ so that $$d𝐱=d𝐫+dq𝐞_3+qd𝐞_3$$ $$=\sigma _1𝐞_1+\sigma _2𝐞_2+\sigma _3𝐞_3.$$ $`(6)`$ To avoid unneccessary complexity, consider an axially symmetric surface given by the set of points that satisfy the surface part of eq. (5), $$𝐫(\rho ,\varphi )=\rho \mathrm{cos}\varphi 𝐢+\rho \mathrm{sin}\varphi 𝐣+S(\rho )𝐤.$$ $`(7)`$ In terms of a local basis, points near the surface can be written either in terms of eq.(5) or in terms of a local set of unit vectors as (subscripts on $`S`$ denote differentiation) $$𝐱(\rho ,\varphi ,q)=𝐫(\rho ,\varphi )+q𝐞_3=\frac{1}{Z}\left[(\rho +SS_\rho )𝐞_1+(S\rho S_\rho +q)𝐞_3\right]$$ $`(8)`$ with $$Z=\sqrt{1+S_{\rho }^{}{}_{}{}^{2}},$$ $`(9)`$ $$𝐞_1=\frac{1}{Z}\left[\mathrm{cos}\varphi 𝐢+\mathrm{sin}\varphi 𝐣+S_\rho 𝐤\right],$$ $`(10)`$ $$𝐞_2=\mathrm{sin}\varphi 𝐢+\mathrm{cos}\varphi 𝐣,$$ $`(11)`$ and $$𝐞_3=\frac{1}{Z}\left[S_\rho \mathrm{cos}\varphi 𝐢S_\rho \mathrm{sin}\varphi 𝐣+𝐤\right].$$ $`(12)`$ The one forms for this system can be read off from $$d𝐱=d\rho Z\left[1\frac{qS_{\rho \rho }}{Z^3}\right]𝐞_1+d\varphi \rho \left[1\frac{qS_\rho }{Z\rho }\right]𝐞_2+dq𝐞_3\sigma _1𝐞_1+\sigma _2𝐞_2+\sigma _3𝐞_3.$$ $`(13)`$ It is possible to write eq.(4) in this geometry, but there is little to be gained by writing the general expression in full detail since our primary interest here is the modification of eq.(4) when the particle is constrained to the surface. A constraint that brings the particle to the surface can be thought of as being effected by a large confining potential everywhere normal to the surface, i.e., a local function of $`q`$ in the $`𝐞_3`$ direction. The usual choice for this term is $$V_n(q)=\frac{1}{2}\omega ^2q^2$$ $`(14)`$ with eventually $`\omega \mathrm{}`$ and $`q0`$. As the particle approaches the surface, we anticipate a decoupling of the wavefunction into tangential and normal degrees of freedom $$\mathrm{\Psi }(\rho ,\varphi ,q)\chi _t(\rho ,\varphi )\chi _n(q).$$ A consistent relation for the norm is obtained by insisting on the condition $$|\mathrm{\Psi }|^2FdSdq=|\chi _t|^2dS|\chi _n|^2dq,$$ $`(15)`$ so that $$\mathrm{\Psi }=\frac{\chi _t\chi _n}{\sqrt{F}}$$ $`(16)`$ with $$F=1+2q_3H+q_3^2K$$ $`(17)`$ and H, K the mean and Gaussian curvatures of the surface. Their explicit forms are given by $$H=\frac{1}{2}\left[\frac{S_\rho }{Z\rho }+\frac{S_{\rho \rho }}{Z^3}\right],$$ $`(18)`$ and $$K=\frac{S_\rho S_{\rho \rho }}{\rho Z^4}.$$ $`(19)`$ The ansatz of eq.(16) can be inserted into eq.(4) to give the following relations for the tangential and normal functions in the $`q0`$ limit: $$\frac{1}{2}\left[Z^2\left(\frac{^2\chi _t}{\rho ^2}+\frac{1}{\rho }\frac{\chi _t}{\rho }\right)\chi _n+\frac{1}{\rho ^2}\frac{^2\chi _t}{\varphi ^2}\chi _n+(H^2K)\chi _t\chi _n+\frac{^2\chi _n}{q^2}\chi _tZ^4S_\rho S_{\rho \rho }\frac{\chi _t}{\rho }\chi _n\right]$$ $$ie\left[\frac{A_1}{Z}\frac{\chi _t}{\rho }\chi _n+\frac{A_2}{\rho }\frac{\chi _t}{\varphi }\chi _n+A_3\frac{\chi _n}{q}\chi _tA_3H\chi _t\chi _n\right]$$ $$+\frac{e^2}{2}\left[A_{1}^{}{}_{}{}^{2}+A_{2}^{}{}_{}{}^{2}+A_{3}^{}{}_{}{}^{2}\right]\chi _t\chi _n+V_n(q)\chi _t\chi _n=i\mathrm{}\frac{}{t}\chi _t\chi _n$$ $`(20)`$ The $`(H^2K)`$ term appearing in the first grouping has been discussed elsewhere and will not prove important for what follows. It is interesting to note that a coupling of $`A_3`$ to the mean curvature $`H`$ obtains even if a limiting procedure does not occur; it originates from a term proportional to $`\frac{F}{q}`$ and is present only if there is variation of the unit normal $`𝐞_3`$. Should the $`q0`$ condition not be imposed, higher order curvature terms would be present in eq.(20). As an admittedly unphysical example of how the curvature term can produce a nontrivial consequence, consider a situation with $`A_1=A_2=0`$, and $`A_3,H`$ both functions of $`\rho `$ only. Next for $$V_n(q)>>A_3\frac{}{q}ln[\chi _n(q)]$$ $`(21)`$ a separable pair of equations obtain, $$\frac{1}{2}\left[Z^2\left(\frac{^2\chi _t}{\rho ^2}+\frac{1}{\rho }\frac{\chi _t}{\rho }\right)\frac{m^2}{\rho ^2}\chi _tZ^4S_\rho S_{\rho \rho }\frac{\chi _t}{\rho }+(H^2K)\chi _t\right]$$ $$+ieA_3H\chi _t+\frac{e^2}{2}A_{3}^{}{}_{}{}^{2}\chi _t=E_t\chi _t$$ $`(22)`$ and $$\frac{1}{2}\frac{^2\chi _n}{q^2}+V_n(q)\chi _n=E_q\chi _n$$ $`(23)`$ with $`m`$ an aziumthal quantum number. Suppose that the product of $`A_3`$ and $`H`$ over some region $`\mathrm{\Gamma }`$ is approximately constant and zero elsewhere. Then for whatever $`\chi _t`$ results from the solving the tangential equation, the solution in $`\mathrm{\Gamma }`$ would become $$\chi _t\chi _t\mathrm{exp}[\pm e|A_3Ht|].$$ $`(24)`$ Eq. (24) illustrates a peculiar situation wherein the sign of a term depending on curvature can have either dissapative or pathological behavior. Cleary this situation is not expected to persist for physically realizable systems, but it serves to show that peculiar results can obtain when studying constrained quantum systems. In this letter, differential forms were used to derive the Schrodinger equation for a three dimensional particle constrained to a two dimensional cylindrically symmetric surface in the presence of a static vector potential. For arbitrary choices of geometry and field configuration, it is difficult to find closed form solutions of eq.(20). Nevertheless, the above discussion indicates that the interplay between surface geometry and applied fields may be important. It is worth reemphasizing that even for finite $`q`$ there is nontrivial coupling of $`A`$ to surface terms through differentiations of eq.(17). This is a manifestation of a general result; imposing constraints in three dimensional space to restrict particle dynamics to a two dimensional space gives different results than an $`a\mathrm{𝑝𝑟𝑖𝑜𝑟𝑖}`$ two dimensional model. Acknowledgments The author would like to acknowledge useful discussions with Ray O’Neal and Lonnie Mott. REFERENCES 1. R. Ashoori, $`\mathrm{𝑁𝑎𝑡𝑢𝑟𝑒}`$, vol. 378, pp. 413-419, 1998. 2. J. R. Chelilowsky and S. G. Louie, eds. $`\mathrm{𝑄𝑢𝑎𝑛𝑡𝑢𝑚}\mathrm{𝑇ℎ𝑒𝑜𝑟𝑦}\mathrm{𝑜𝑓}\mathrm{𝑅𝑒𝑎𝑙}\mathrm{𝑀𝑎𝑡𝑒𝑟𝑖𝑎𝑙𝑠}`$, (Kluwer Academic Publishers, Massachusetts, 1996). 3. D. Pfannkuche, R. Gerhardts, P. Maksym andV. Gudmundsson, Phys. B $`\mathrm{𝟏𝟖𝟗}`$, 6, (1993). 4. R. C. T. da Costa, Phys. Rev. A $`\mathrm{𝟐𝟑}`$, 1982 (1981). 5. R. C. T. da Costa, Phys. Rev. A $`\mathrm{𝟐𝟓}`$, 2893 (1982). 6. M. Burgess and B. Jensen, Phys. Rev. A $`\mathrm{𝟒𝟖}`$, 1861 (1993). 7. S. Matsutani, J. Phys. Soc. Japan 61, $`\mathrm{𝟓𝟓}`$, (1992). 8. S. Matsutani and H. Tsuru, J. Phys. Soc. Japan, $`\mathrm{𝟔𝟎}`$, 3640 (1991). 9. M. Encinosa and B. Etemadi, Physical Review A $`\mathrm{𝟓𝟖}`$, 77 (1998). 10. B. Skarsgard, $`\mathrm{𝐺𝑒𝑜𝑚𝑒𝑡𝑟𝑦},\mathrm{𝑃𝑎𝑟𝑡𝑖𝑐𝑙𝑒𝑠}\mathrm{𝑎𝑛𝑑}\mathrm{𝐺𝑎𝑢𝑔𝑒}\mathrm{𝐹𝑖𝑒𝑙𝑑𝑠}`$ (Springer Verlag, New York, 1998). 11. H. Flanders; $`\mathrm{𝐷𝑖𝑓𝑓𝑒𝑟𝑒𝑛𝑡𝑖𝑎𝑙}\mathrm{𝐹𝑜𝑟𝑚𝑠}\mathrm{𝑤𝑖𝑡ℎ}\mathrm{𝐴𝑝𝑝𝑙𝑖𝑐𝑎𝑡𝑖𝑜𝑛𝑠}\mathrm{𝑡𝑜}\mathrm{𝑡ℎ𝑒}\mathrm{𝑃ℎ𝑦𝑠𝑖𝑐𝑎𝑙}\mathrm{𝑆𝑐𝑖𝑒𝑛𝑐𝑒𝑠}`$ (Dover Books, New York, 1989). 12. M. Encinosa and B. Etemadi, Physica B, 266 , 361 (1999).
no-problem/9908/cond-mat9908484.html
ar5iv
text
# Critical Quantum Chaos in 2⁢𝐷 Disordered Systems with Spin-Orbit Coupling ## Abstract We examine the validity of the recently proposed semi-Poisson level spacing distribution function $`P(S)`$, which characterizes ‘critical quantum chaos’, in $`2D`$ disordered systems with spin-orbit coupling. At the Anderson transition we show that the semi-Poisson $`P(S)`$ can describe closely the critical distribution obtained with averaged boundary conditions, over Dirichlet in one direction with periodic in the other and Dirichlet in both directions. We also obtain a sub-Poisson linear number variance $`\mathrm{\Sigma }_2(E)\chi _0+\chi E`$, with asymptotic value $`\chi 0.07`$. The obtained critical statistics, intermediate between Wigner and Poisson, is relevant for disordered systems and chaotic models. In mesoscopic physics the effect of disorder on the electron propagation leads to the zero-temperature quantum Anderson metal-insulator transition, which arises from the competition between quantum tunelling and interference, as a function of disorder. For weak disorder the electrons diffuse, due to quantum tunelling, and the system is metallic with correlated chaotic energy levels and ‘level-repulsion’ described by Wigner statistics . In the case of strong disorder the electrons localize in random positions, due to quantum interference, and the system becomes insulating, having non-chaotic completely uncorrelated (random) energy levels which show ‘level attraction’ and obey ordinary Poisson statistics. In order to see the metal-insulator transition high enough space dimensionality (usually greater than $`2`$) is required and at the critical point, which corresponds to an intermediate value of disorder, the level statistics changes from Wigner to Poisson . The critical electrons are neither extended nor localized and it is believed that a new universal critical statistics, intermediate between Wigner and Poisson, should apply . We aim to address the question of the critical statictics in two dimensions ($`2D`$), where a metal-insulator transition occurs in the presence of spin-orbit coupling . The stationary energy levels of $`2D`$ quantum billiards (e.g. in the form of the stadium), with zero potential inside and infinity outside, can also display quantum chaotic behavior . The analogies in the level statistical description bring together the two fields of mesoscopic physics and quantum chaos and have been exploited in the past for understanding important phenomena in both areas. In this respect, Wigner statistics was originally conjectured to apply for quantum systems with chaotic classical dynamics, since the levels resemble the eigenvalues found in appropriate random matrix ensembles, introduced long ago . On the other hand, integrable systems correspond to Poisson statistics having completely uncorrelated (random) eigenvalues. The key question is again what happens at criticality, between chaos and integrability, similarly to the transition between metal and insulator. Recently, a new distribution was proposed to describe critical levels statistics , which contains both Wigner and Poisson features, as the main theme of what is called ‘critical quantum chaos’. This intermediate distribution can be derived from a short range plasma model and was realized in pseudo-integrable systems, such as the classically non-integrable but of zero metric entropy rational triangle billiards , and corresponds to other solvable models . In disordered systems the intermediate distribution, named semi-Poisson, was shown to characterize critical states at the $`3D`$ metal-insulator transition and the energy levels of few electrons in the presence of interactions . FIG. 1. Energy levels $`E`$ versus disorder $`W`$, for a $`2D`$ disordered system with spin-orbit coupling. The metal-insulator critical point $`W_c=8.55`$, marked with a broken line, separates chaotic levels (on the left) with ‘level-repulsion’ and ‘spectral rigidity’ described by Wigner statistics, from non-chaotic levels (on the right) with ‘level attraction’ and ‘spectral randomness’ described by Poisson statistics. In this paper we address what happens at the metal-insulator transition (broken line). In Fig.1 the energy levels $`E`$ obtained from our $`2D`$ disordered system are displayed as a function of disorder $`W`$. The behavior of the levels is seen clearly to change at the critical point of the metal-insulator transition $`W_c`$, which separates chaotic levels (on the left) from non-chaotic levels (on the right). The chaotic levels are more regular (correlated) than the non-chaotic levels, which are uncorrelated (random). The ‘level-repulsion’ effect can be seen on the chaotic levels and the ‘level-attraction’ on the non-chaotic levels, where degeneracies exist. The displayed levels in Fig. 1 are obtained for one random configuration from a system of linear size $`L=30`$, by keeping all the energies at the middle of the spectrum, making the density of states constant at every $`W`$. The overall ‘spectral rigidity’, seen in the metallic chaotic levels, can be contrasted with the ‘spectral randomness’ of the insulating non-chaotic levels. The surprising result, also visualized in this figure, is that the chaotic levels appear more ‘regular’ than the non-chaotic ones. The question addressed in this paper is: ‘what is the level statistics at the critical broken line of Fig. 1?’ This study is done in connection to the scenario of ‘critical quantum chaos’, which is summarized in a semi-Poisson level spacing distribution function $`P(S)`$ and a sub-Poisson linear number variance $`\mathrm{\Sigma }_2(E)`$, which measures level-fluctuations in a given energy window $`E`$. Moreover, in order to examine the validity of the semi-Poisson $`P(S)`$ different boundary conditions (BC) must be considered, in the spirit of recent important findings . The fact that the critical level fluctuations are at the same time scale-invariant (size-independent) and dependent on BC was explained by invoking the concept of the critical conductance . The influence of boundary conditions on the critical level statistics, without affecting the critical disorder, has also been demonstrated for a different critical $`2D`$ model . The main features of ‘critical quantum chaos’ for systems characterized by the universality index $`\beta =1,2,4`$ are summarized in: (i) The semi-Poisson $`P(S)`$ level spacing distribution which shows Wigner-like repulsion $`S^\beta `$ at small spacings $`S<<1`$ and is exponential, Poisson-like, $`\mathrm{exp}((\beta +1)S)`$ at large spacings $`S>>1`$, overall described by the scale-invariant normalized semi-Poisson curve $$P(S)=AS^\beta \mathrm{exp}((\beta +1)S),$$ (1) with the constant values $`A=4`$, $`27/2`$ and $`3125/24`$ obtained from normalization, respectively. The spacing distribution $`P(S)`$, by applying a ‘level-unfolding’ procedure which keeps the level-density constant, corresponds to $`S=1`$. (ii) The sub-Poisson number variance, which defines the level number fluctuations in an energy window $`E`$, with the mean number proportional to $`E`$ after ‘unfolding’, according to this scenario is $$\mathrm{\Sigma }_2(E)\chi _0+\chi E,$$ (2) defining the level compressibility $`\chi `$. The value of $`\chi `$ ranges between $`0`$ (chaos) and $`1`$ (integrability) and was related to the multifractality of the critical wavefunctions . The considered disordered system displays a transition in $`2D`$ with energy levels which obey Wigner statistics for the metal (with $`\beta =4`$) and Poisson statistics for the insulator (see Fig. 1). At criticality, where one expects ‘critical quantum chaos’ to apply, numerical work suggested level-repulsion in $`3D`$ for small $`S`$ , also later shown in $`2D`$ . In order to study carefully the level fluctuations in the critical region it is important to identify the crucial role of BC . We find that for the three considered kinds of BC the critical distribution function shows level repulsion at small spacings and is Poisson-like at large spacings. However, when considering an averaged distribution over the cases: 1) Dirichlet BC in both directions and 2) periodic BC in one direction and Dirichlet in the other, the obtained distribution is seen to be remarkably close to the scale-invariant semi-Poisson curve of Eq. (1), appropriate for $`\beta =4`$ (see Fig. 3 below). The theoretical framework to study the Anderson transition can classify disordered systems into three universality classes, depending on whether the Hamiltonian preserves the time-reversal invariance or the rotational invariance, in direct analogy with the random matrix theory description of quantum chaotic systems . Zero spin-orbit corresponds to the orthogonal universality class ($`\beta =1`$) and finite spin-orbit breaks the rotational invariance so that one obtains the symplectic universality class ($`\beta =4`$). In our calculations we consider a two-dimensional disordered system, with spin-orbit coupling for spin-$`\frac{1}{2}`$ particles, described by the Hamiltonian $$=\underset{i,\sigma }{}ϵ_ic_{i,\sigma }^+c_{i,\sigma }+\underset{(i,j)}{}\underset{(\sigma ,\sigma ^{^{}})}{}V_{i,j;\sigma ,\sigma ^{^{}}}c_{i,\sigma }^+c_{j,\sigma ^{^{}}},$$ (3) where $`i`$ labels the $`L^2`$ square lattice sites and $`\sigma =\pm 1/2`$ is the spin index on each site. The second sum is taken over all nearest neighbour lattice pairs $`(i,j)`$ and the random on site potential $`ϵ_i`$ is a spin independent uniformly distributed random variable, chosen from a probability distribution of width $`W`$. In this case the nearest neighbour hoppings $`V_{i,j}`$ are random $`2\times 2`$ matrices describing spin rotation, due to spin-orbit, on every lattice bond $`(i,j)`$. In the spinor space they are represented by $$V_{i,j}=\left(\begin{array}{cc}1+i\mu V^z\hfill & \mu V^y+i\mu V^x\hfill \\ \mu V^y+i\mu V^x\hfill & 1i\mu V^z\hfill \end{array}\right)_{ij},$$ (4) where $`\mu `$ denotes the spin-orbit coupling and the $`V^x,V^y`$ and $`V^z`$, defined for every bond $`(i,j)`$, are real and independent random variables chosen from a uniform probability distribution on the interval $`[\frac{1}{2},+\frac{1}{2}]`$. For the rest the spin-orbit strength is fixed to $`\mu =2`$ and the disorder is chosen to lie exactly at the critical point $`W_c=8.55`$ . We compute the eigenvalues from Eq. (3) by diagonalizing numerically the corresponding Hamiltonian matrices for large square lattices. The statistical analysis of energy levels must be done on a constant density of states using an ‘unfolding procedure’. In order to achieve the level unfolding for the considered disordered system it is sufficient to obtain the average of the integrated density of states $`𝒩`$, locally at $`E`$, by repeating many times the disorder configuration creating a statistical ensemble. Then the ‘raw’ spacings $`\mathrm{\Delta }_i=E_iE_{i1}`$ are replaced by the ‘unfolded’ new ones $`S_i=𝒩_{av}(E_i)𝒩_{av}(E_{i1})(E_iE_{i1})\frac{𝒩_{av}(E)}{E}=(E_iE_{i1})/\mathrm{\Delta }`$, where $`\mathrm{\Delta }`$ is the local mean spacing around $`E_i`$ or equivalently the inverse density of states of the raw data. In the numerical calculations we considered eigenvalues within the energy window $`[2,2]`$ performing $`2700,1200,675`$ and $`794`$ random configuration runs in each case, with $`L=20,30,40`$ and $`60`$, respectively. The total number of eigenvalues from all random configurations for each BC is about $`400,000`$ for $`L=20,30,40`$ and $`1,000,000`$ for $`L=60`$. These ‘raw’ data were ‘unfolded’ in the described way. FIG. 2. This figure shows the variation of the critical $`P(S)`$ for three choices of BC for a system of linear size $`L`$. The mean distribution over the three cases is shown by the continuous black line. The Wigner distribution (dotted line) is also shown. The obtained $`P(S)`$ at criticality is shown in Fig. 2 for the three different choices of BC. The computed curves are, clearly, very different, in agreement with the corresponding $`3D`$ results for $`\beta =1`$ . They are very different from the Wigner or Poisson curves while their average, over the three BC, cannot fit to the semi-Poisson, either. However, the average over Dirichlet (hard wall) in both directions $`(00)`$ and periodic in one direction with Dirichlet in the other $`(10)`$, which is displayed in Fig. 3 for various system sizes, gives a distribution very well described by the semi-Poisson curve of Eq. 1. This is the most important result of the paper and shows the validity of the semi-Poisson for the chosen specific average over BC at criticality. The obtained semi-Poisson is also in agreement with recent results for the critical $`P(S)`$ at the metal-insulator transition in $`3D`$ disordered systems, where the semi-Poisson was obtained by averaging over all possible combinations of BC . FIG. 3. The mean $`P(S)`$ distribution of the (10) and (00) combinations of BC for several system sizes $`L`$ is shown to follow closely the semi-Poisson distribution $`P(S)=(3125/24)S^4\mathrm{exp}(5S)`$ (Eq. (1) for $`\beta =4`$) (black line). The Wigner (dotted line) and the Poisson (dashed line) are also plotted. The longer, in the $`E`$-range, critical spectral fluctuations are described by a linear number variance $`\mathrm{\Sigma }_2(E)\chi _0+\chi E`$, with the compressibility $`\chi `$ related to the critical wavefunction dimension $`D_2^\psi `$ and the space dimension $`d`$ via $`\chi =(1/2)(1D_2^\psi /d)`$ . In the considered model previous studies gave $`D_2^\psi 1.63`$ . For a rather small energy window $`E`$ (see Fig. 4) the level number variance is shown to be linear with level compressibility $`\chi `$ which varies with the chosen BC. However, when the energy window $`E`$ increases the result becomes independent of BC, as was already shown in $`3D`$ . The obtained asymptotic level compressibility in this case becomes $`\mathrm{\Sigma }_2(E)/E\chi 0.07`$ which leads to a $`D_2^\psi `$, rather close to the expected value according to the previous formula. The obtained value of $`\chi `$ in $`2D`$ should be contrasted with the higher $`3D`$ asymptotic value $`\chi 0.27`$ . The main result from our calculations, done on a square random lattice and not on a peculiarly shaped non-random billiard, is the validity of the semi-Poisson statistics at the metal-insulator transition in $`2D`$ disordered systems with spin-orbit coupling. However, the semi-Poisson $`P(S)`$ is obtained for the averaged distribution over two specific BC. It must be pointed out that only the main part of the distribution agrees surprisingly well with the analytical result. For large $`S`$, where the dependence on BC should become less important, we have not suceeded to describe its tails, possibly due to their exponentially small nature. In this case the appropriate statistical measure becomes the number variance since longer range level correlations are needed. We find a linear number variance $`\mathrm{\Sigma }_2(E)\chi E`$ which becomes independent of the BC choice. The obtained $`\chi `$ is close to the expected value from the formula via $`D_2^\psi `$. FIG. 4. The number variance $`\mathrm{\Sigma }_2(E)`$ vs the energy window $`E`$ for the two BC (00) and (10). The straight lines fit the data giving slopes corresponding to non-asymptotic $`\chi `$ values. The Wigner (dotted line) and the Poisson (dashed line) are also displayed for comparison. In conclusion, we have shown the validity of the semi-Poisson level statistics at the critical point of the metal-insulator transition with $`\beta =4`$ in $`2D`$. The semi-Poisson curve is shown to describe very well the main part of the computed distribution for a specific average over BC and is similar to recent results for critical disordered systems and weakly chaotic quantum systems. Our calculations, on one hand, can give a justification for the averaging over boundary conditions recently shown to lead to the semi-Poisson statistics at the mobility edge . On the other hand, suggest that such an average might be related to the bandwidth distributions, by repeating periodically the square, as it was recently shown for a non-random one-dimensional critical quasi-periodic model . Clearly, more work is needed to examine the validity of the critical semi-Poisson distribution, also for $`3D`$ disordered systems with spin-orbit coupling and systems in the presence of a magnetic field ($`\beta =2`$). This work was supported by a TMR network of EU and a Greek-Chineese collaboration G.S.S.T. Useful discussions with J.-L. Pichard and G. Montambaux are gratefully aknowledged. FIG. 5. $`\mathrm{\Sigma }_2(E)/E`$ versus $`E`$ for a much broader range of $`E`$ where the independence on BC is seen. The asymptotic value approximates $`\chi 0.07`$.
no-problem/9908/cond-mat9908088.html
ar5iv
text
# 1. Introduction ## 1. Introduction The recent, rather controversial discussions about magnetic ordering in quasicrystals, in particular for alloys containing rare earth elements , has revived the interest in simple models of magnetic ordering in quasicrystals and their properties. The influence of an aperiodic order on magnetic phase transitions can be understood by heuristic scaling arguments resulting in the Harris-Luck criterion for the relevance of aperiodicity. For one-dimensional aperiodic Ising quantum chains, equivalent to two-dimensional layered Ising models with one-dimensional aperiodicity, exact real-space renormalization techniques can be used that prove the heuristic criterion for a large class of models based on substitution systems . For the classical Ising model defined on two-dimensional quasiperiodic cut-and-project tilings, the criterion predicts that the critical behaviour is the same as for periodic lattices, hence these models belong to the Onsager universality class. This has unanimously been corroborated by results obtained from various approaches , and has been commonly accepted. Nevertheless, it is interesting to study these models in more detail, and in this short note we present results for periodic approximants for the Penrose and the Ammann-Beenker tiling. The zero-field Ising model on periodic approximants can be treated by methods that allow us to compute expansion coefficients and partition function zeros for the infinite periodic tiling, hence we do not have to deal with finite-size corrections as one usually accounts if only finite patches are considered. Instead, we can study the dependence on the size of the unit cell. From the partition function zeros, we obtain, in principle, the exact critical temperatures for the periodic approximants. In this way, we are able to derive precise estimates for the critical temperatures of the quasiperiodic models. For the high-temperature series of the free energy, we investigate the fluctuations in the sequence of expansion coefficients observed in , using the exact values for the critical temperatures of the periodic approximants. ## 2. The Ising model on periodic approximants We place an “Ising spin” $`\sigma _i=\pm 1`$ on each vertex $`i`$ of the periodic approximant. A configuration $`\sigma =(\sigma _1,\sigma _2,\mathrm{},\sigma _n)`$ is assigned the energy $$E(\sigma )=J\underset{i,j}{}\sigma _i\sigma _j$$ (1) where $`J>0`$ is the ferromagnetic exchange coupling, which we choose identical for all pairs of neighbouring spins, and we sum over all neighbouring pairs of spins $`i,j`$ at vertices $`i`$ and $`j`$. Two vertices are neighbours if they are connected by an edge of the tiling. There exist several variants for the solution of the zero-field Ising model; the method we use here was initially formulated by Kac and Ward for the square lattice Ising model . The Kac-Ward determinant and its application to periodic approximants was described in detail in our recent paper . Let us just mention briefly that the partition function $`Z(\beta )=_\sigma \mathrm{exp}[\beta E(\sigma )]`$ for a zero-field Ising model on an arbitrary planar graph, where $`\beta =1/k_BT`$ denotes the inverse temperature, is expressed as the square root of a determinant of an infinite matrix $`K(w)`$, where $`w=\mathrm{tanh}(\beta J)`$ . For a periodic approximant, this matrix is cyclic. Thus the problem can, by Fourier transform, be reduced to the computation of a determinant of a finite matrix $`\stackrel{~}{K}(w,\phi _1,\phi _2)`$ that depends on two additional parameters $`\phi _1,\phi _2[0,2\pi )`$ which play the role of Bloch phases. The size of the matrix $`\stackrel{~}{K}(w,\phi _1,\phi _2)`$ is given by the number of oriented edges in the unit cell, hence by $`4N`$ for the Penrose and the Ammann-Beenker approximants, where $`N`$ denotes the number of spins in the unit cell. Apart from a factor $`1/\beta `$, the free energy per vertex in the thermodynamic limit $`f(w)=lim_M\mathrm{}[\mathrm{log}Z(\beta )/M]`$ is given as $$f(w)=\frac{1}{2N}\underset{0}{\overset{2\pi }{}}\underset{0}{\overset{2\pi }{}}\frac{d\phi _1}{2\pi }\frac{d\phi _2}{2\pi }\mathrm{log}det\stackrel{~}{K}(w,\phi _1,\phi _2)$$ (2) where $`M`$ denotes the the number of spins in a finite patch approximating the infinite system. For large approximants, it is expendable, and not very profitable, to compute the determinant analytically as a function of the parameters $`w`$, $`\phi _1`$, and $`\phi _2`$. Instead, we rather calculate the partition function zeros or the coefficients of the high-temperature expansion of the free energy by exploiting the fact that the matrix $`\stackrel{~}{K}`$ has the form $$\stackrel{~}{K}(w,\phi _1,\phi _2)=I+w\stackrel{~}{L}(\phi _1,\phi _2)$$ (3) where $`I`$ denotes the unit matrix. In particular, the matrix $`\stackrel{~}{L}`$ does not depend on the variable $`w`$. This makes it possible to compute the partition function zeros for given values of $`\phi _1`$ and $`\phi _2`$ from the eigenvalues of the matrix $`\stackrel{~}{L}(\phi _1,\phi _2)`$, and the complete set of partition function zeros is just the accumulation of the zeros for all $`\phi _1,\phi _2[0,2\pi )`$. The coefficients of the high-temperature expansion of the free energy $$f(w)=\underset{n=2}{\overset{\mathrm{}}{}}g_{2n}w^{2n}$$ (4) can be calculated from traces of powers of $`\stackrel{~}{L}`$ by $$g_{2n}=\frac{1}{4nN}\underset{0}{\overset{2\pi }{}}\underset{0}{\overset{2\pi }{}}\frac{d\phi _1}{2\pi }\frac{d\phi _2}{2\pi }\text{tr}[\stackrel{~}{L}^{2n}(\phi _1,\phi _2)]$$ (5) which follows by expanding the logarithm in Eq. (2). ## 3. Partition function zeros In what follows, we adopt the usual convention and represent the partition function zeros in the variable $`z=(1+w)/(1w)=\mathrm{exp}(2\beta J)`$. For an infinite system, the zeros of the partition function in the complex $`z`$ plane accumulate on lines or areas that separate different regions of analyticity of the free energy, hence different phases of the model. In particular, partition function zeros on the real axis correspond to phase transition points, and thus determine the critical temperature. The behaviour of zeros around these points is governed by the corresponding critical exponents . From the present approach, it is obvious that the zeros themselves are parametrized by the two angles $`\phi _1`$ and $`\phi _2`$, thus we expect that they generically fill areas in the complex plane. In the square lattice case, however, the zeros fall on two circles with radius $`\sqrt{2}`$, centred at $`z=\pm 1`$; similar results are found for other simple planar lattices. In order to illustrate the behaviour for the periodic approximants, we show the zero patterns for the third approximant of the Penrose tiling with $`N=76`$ spins per unit cell in Fig. 1, and the result for the second approximant of the Ammann-Beenker tiling with $`N=41`$ in Fig. 2. Here, the zeros were computed for angles $`\phi _1,\phi _2\{m\pi /20m=0,1,2,\mathrm{},39\}`$. Clearly, the patterns are more complicated than the two circles found for the square lattice, in particular further away from the positive real axis. We find two zeros on the positive real axis, one for $`\phi _1=\phi _2=0`$ and the other for $`\phi _1=\phi _2=\pi `$, corresponding to the ferromagnetic and antiferromagnetic critical points, respectively, which have the same properties due to the bipartiteness of the tilings. In the vicinity of the critical points, the zero patterns are very well described by segments of circles that orthogonally intersect the real axis, showing that the corresponding critical exponent $`\alpha `$ stays the same as for the square lattice, i.e. $`\alpha =0`$, corresponding to a logarithmic singularity of the specific heat. The zeros on the real axis determine the critical temperatures for the periodic approximants, the results for the ferromagnetic critical point are shown in Tables 1 and 2. The values are in agreement with results of recent Monte Carlo simulations using the invaded cluster algorithm . Interestingly, the values for the Ammann-Beenker approximants in Table 2 appear to converge much faster than those for the Penrose case in Table 1. This might be related to the observation that the mean coordination numbers , in particular the number of next-nearest neighbours, also converge faster for the Ammann-Beenker case, thus the periodic approximants of the Ammann-Beenker tiling are, in this sense, “closer” to the infinite quasiperiodic case than those of the Penrose tiling. ## 4. High-temperature expansion In principle, the expansion coefficients of the high-temperature series can be calculated exactly for quasiperiodic cut-and-project tilings in the framework of the projection method. However, the number of graphs contributing to the expansion grows tremendously, thus limiting the applicability of this method. Recently , we presented the leading terms, up to 18th order in the expansion variable $`w=\mathrm{tanh}(\beta J)`$, of the high-temperature series (4) for the free energy $`f(w)`$ of the Ising model on the Penrose and the Ammann-Beenker tiling. The series coefficients contain information about the critical behaviour of the model. Assuming that the free energy $`f(w)`$ at the critical point $`w_c`$ shows a power-law singularity $`f(w)(1w^2/w_c^2)^\kappa `$, the ratios of successive coefficients behave as $$\frac{g_{2n}}{g_{2n2}}\frac{1}{w_c^2}\left(1\frac{\kappa +1}{n}\right)$$ (6) for large $`n`$, compare Eq. (4). Thus, plotting the ratio $`g_{2n}/g_{2n2}`$ as a function of $`1/n`$, the data points should, for sufficiently large values of $`n`$, lie on a straight line. From its intercept at $`1/n=0`$ and its slope, we can, in principle, extract the values of the critical temperature $`w_c`$ and the critical exponent $`\kappa =2\alpha =\nu d`$. Here, $`d=2`$ denotes the spatial dimension, $`\nu `$ is the critical exponent of the correlation length, which is $`\nu =1`$ for the Onsager universality class, and the last equality is due to scaling. As shown in , this plot reveals large fluctuations, in particular for the smallest periodic approximants, while it behaves very nicely for simple lattices as, for instance, the square lattice. These fluctuations rendered impossible any reasonable estimate of the critical temperature or the critical exponent from the series expansion, at least in this way. In order to get some more insight into the nature and the reasons of the fluctuations, we extended the series expansions for the periodic approximants considerably. In Fig. 3, we plot the differences $$r_{2n}=\frac{n3}{nw_c^2}\frac{g_{2n}}{g_{2n2}}$$ (7) between the ratios of the expansion coefficients, up to order $`2n=128`$, and their expected asymptotic behaviour (6), with $`\kappa =2`$, for the smallest approximants of the Penrose tiling. Here, we use the critical temperatures $`w_c`$ given in Table 1. Apparently, the ratios $`r_{2n}`$ for the smallest approximant with $`N=11`$ spins per unit cell fluctuate strongly for $`20\stackrel{<}{}2n\stackrel{<}{}60`$. For larger values of $`2n`$, the deviations from the asymptotic behaviour are small, similar to the square lattice case. The second approximant with $`N=29`$ shows a strikingly similar region with strong fluctuations, but at larger orders, starting around $`2n\stackrel{>}{}50`$ and waning just around the maximal order $`2n=128`$ in Fig. 3. For the third approximant, no comparable region of particularly strong fluctuations is observed, but we might speculate that the order where these fluctuations occur grows roughly with the number of spins in the unit cell. In this case one has to go to higher orders to observe the effect. One might expect that for orders clearly beyond the number of oriented edges in the unit cell, i.e., for $`n4N`$, the deviations from the asymptotic behaviour (6) are small. For larger approximants, the expansion coefficients $`g_{2n}`$ rapidly approach those of the quasiperiodic tilings . Thus, we may draw the conclusion that the data for the periodic approximants show cross-over phenomena between the characteristic behaviour for the quasiperiodic model and the simple square-lattice case, although for the cases considered here the asymptotic behaviours should coincide. It seems reasonable that the cross-over occurs approximately at the size of the unit cell. ## 5. Concluding remarks The Ising model defined on quasiperiodic tilings such as the Ammann-Beenker and the Penrose tiling belongs, according to scaling arguments, to the same universality class as the model on the square lattice. This is in accordance with our results on the partition function zeros and the high-temperature expansions for the corresponding periodic approximants. Nevertheless, the quasiperiodic systems show interesting features, and a better understanding of the cross-over phenomena in the high-temperature expansions of the periodic approximants is desirable. ## Acknowledgements U. G. thanks M. Baake and O. Redner for discussions and helpful comments. This work has been supported by Deutsche Forschungsgemeinschaft (DFG).
no-problem/9908/quant-ph9908001.html
ar5iv
text
# General Strategies for Discrimination of Quantum States ## Abstract We derive general discrimination of quantum states chosen from a certain set, given initial $`M`$ copies of each state, and obtain the matrix inequality, which describe the bound between the maximum probability of correctly determining and that of error. The former works are special cases of our results. PACS numbers: 03.67.-a, 03.65.Bz, 89.70.+c It is determined by the principle of quantum mechanics that we can’t perfectly discriminate or clone an arbitrary,unknown quantum states. But this important result doesn’t prohibit discrimination and cloning strategies which has a limited success. Numerous projects has been made to this subject by many author. Discrimination of states has a close connection with quantum measurement. The general approach of quantum measurement theory has been consider a quantum system whose unknown state belongs to a finite, known set and devise the measurement which yields the most information about initial state, where the figure of merit is the probability of a correct result or the mutual information . We have known that a set of orthogonal quantum states can be discriminated perfectly. Approximate state determination of a set of non-orthogonal states is also possible. Helstrom has found the absolute maximum probability of discriminating between two states, which is instead given by the well-known Helstrom limit . Non-orthogonal quantum states can also,with some probability, be discriminated without approximation. Ivanovic , Dieks and Peres have showed that it is possible to discriminate exactly ,which means with zero error probability, between a pair of non-orthogonal states and they derive the maximum probability of success called IDP limit. The IDP limit is not the absolute maximum of the discrimination probability, but is rather the maximum subject to the constraint that the measurement never give error results. Chelfes and Barnett \[6-10\] extent their results to the constraint of $`M`$ initial copies of states and also they consider the results for $`n`$ states. They found that $`n`$ non-orthogonal states can be probabilistic discriminate without error if and only if they are linearly independent. Their works connect the approximate and exact discrimination and give an inequality that describe the relation between the probability of correctly and error discriminating. All their works have use an assumption that the probabilities of correct or error for discrimination are equal for different state. More recently, Duan and Guo have found the maximum probabilities of exact discrimination of $`n`$ linearly independent quantum states and for different state the maximum probability may be different. Exact discrimination attempts may give inconclusive results,but we can always know with certainty whether or not the discrimination has been a success. Recently, however, Massar and Popescu and Derka et al have considered the problem of estimating a completely unknown quantum state, given $`M`$ independent realizations. In quantum mechanics, a combination of unitary evolution together with measurements often yields interesting results, such as the quantum programming , the purification of entanglement ,and the teleportation and preparation of quantum states. Duan and Guo used such combination in the field of quantum cloning and showed that the states secretly chosen from a certain set $`S=\{|\mathrm{\Psi }_1,|\mathrm{\Psi }_2,\mathrm{},|\mathrm{\Psi }_n\}`$ can be probabilistically cloned if and only if $`|\mathrm{\Psi }_1`$,$`|\mathrm{\Psi }_2`$,$`\mathrm{}`$,and $`|\mathrm{\Psi }_n`$ are linearly independent. They derive the best cloning efficiencies and also extent their results to the discrimination of a set of states and give the optimal efficiencies for exactly probabilistically discriminating. In this paper we’ll also use such combination and construct general discrimination of the states secretly chosen from a certain set $`S`$, given $`M`$ initial copies of each state. We derive the matrix inequality, which describe the bound between the maximum probability of correctly determining and that of error. We give the most general results about quantum discrimination in a finite set of states and prove that the former works can be derived from our results in different conditions. Consider a set of quantum states $`S=\{|\mathrm{\Psi }_1,|\mathrm{\Psi }_2,\mathrm{},|\mathrm{\Psi }_n\}`$ ,$`S`$ ,where $``$ is a $`n`$-dimension Hilbert space. If there are $`M`$ quantum systems, all of which are prepared in the same states, then we can denote the possible states of the combined system are the $`M`$-fold tensor products $$|\mathrm{\Psi }_i^M=|\mathrm{\Psi }_i_1\mathrm{}|\mathrm{\Psi }_i_M\text{ }$$ (1) Any operation in quantum mechanics can be represented by a unitary evolution together with a measurement. We use a unitary evolution $`\widehat{U}`$ and yield $$\widehat{U}|\mathrm{\Psi }_i_A^M|P_0=\sqrt{\gamma _i}|\phi _i_A|P_0+\underset{j}{}t_{ij}|\phi _j_A|P_0+\underset{j}{}c_{ij}|\phi _j_A|P_1,$$ (2) where $`t_{ii}=0`$, $`\gamma _i0`$, $`t_{ij}0`$, $`\left\{|\phi _i_A,i=1,2,\mathrm{},n\right\}`$ are a set of orthogonal states in Hilbert space $`^M`$, $`\{|P_0,|P_1\}`$ are the orthogonal basis states of the probe system $`P`$, and $`A`$ represent the system of initial states. After the unitary evolution we measure the probe system $`P`$. If we get $`|P_1`$,the discrimination fails and we discard the output, so we call this situation as an inconclusive result, else if we get $`|P_0`$, we then measure system $`A`$ and if get $`|\phi _k_A`$ ,we determine the initial states as $`|\mathrm{\Psi }_k^M`$. We can see the discrimination may be correctly success (when $`k=i`$, the probability is $`P_D^{(i)}=\gamma _i`$) or have errors (when $`ki`$, the all probability is $`P_E^{(i)}=_kt_{ik}^2`$) when the initial state is $`|\mathrm{\Psi }_i^M`$.We can also give the inconclusive probability $`P_I^{\left(i\right)}=1P_D^{(i)}P_E^{(i)}`$. Unitary evolution $`\widehat{U}`$ exist if and only if $$X^{(M)}=(\sqrt{\mathrm{\Gamma }}+T)(\sqrt{\mathrm{\Gamma }}+T^+)+CC^+,$$ (3) where the $`n\times n`$ matrices $`X^{(M)}=\left[\mathrm{\Psi }_i\mathrm{\Psi }_j^M\right]`$, $`\mathrm{\Gamma }=diag(\gamma _1,\gamma _2,\mathrm{}\gamma _n)`$, $`C=\left[c_{ij}\right]`$, $`T=\left[t_{ij}\right]`$, we call $`\mathrm{\Gamma }`$ as correctly discriminating probabilistic matrix and $`T`$ as error discriminating probabilistic matrix. $`CC^+`$ is semipositive definite,that yield $$X^{(M)}(\sqrt{\mathrm{\Gamma }}+T)(\sqrt{\mathrm{\Gamma }}+T^+)0,$$ (4) where $`0`$ means semipositive definite. This inequality give a general bound among the initial states matrix $`X`$, correctly discriminating probabilistic matrix $`\mathrm{\Gamma }`$ and error discriminating probabilistic matrix $`T`$. We don’t make constraint on the initial states, which means we don’t demand that the initial states must be linearly independent. In the following we discuss Inequality.(4) and find many former works about discrimination can be derived from this inequality. Denote $`B=\sqrt{\mathrm{\Gamma }}+T`$,we rewrite inequality.(4) as $$X^{(M)}BB^+0.$$ (5) We begin our discuss with giving a condition that yield $`\left\{|\mathrm{\Psi }_i^M,i=1,2,\mathrm{},n\right\}`$ are linearly independent, that is $$\sqrt{\gamma _i}>\underset{j}{}t_{ij},$$ (6) where $`i=1,2,\mathrm{},n`$. Condition (6) yield $`X^{(M)}`$ is a positive definite matrix , which means$`\left\{|\mathrm{\Psi }_i^M,i=1,2,\mathrm{},n\right\}`$ are linearly independent. We can derive from inequality.(6) that $`P_D^{(i)}>P_E^{(i)}`$, which means the correctly discriminating probability is great than that of error. We consider a special situation that $`T=0`$. To obtain maximum of correctly discriminating probability $`\gamma _i>0`$, the inequality (6) must be satisfied and $`\left\{|\mathrm{\Psi }_i^M,i=1,2,\mathrm{},n\right\}`$ must be linearly independent, which means only linearly independent states can be discriminated with non-zero probability if we demand there are no error existence. Such result has been obtained by Duan and Guo and Chefles and Barnett . We can also give the maximum discriminating probability that is determined by such inequality $$X^{(M)}\mathrm{\Gamma }0.$$ (7) This inequality is just a generalization of the optimal efficiencies for exactly probabilistically discriminating which has been given by Duan and Guo and we also obtain such inequality in . In the following discussion we are only concerned with the discrimination of two states. This situation is the most important and many valuable works has been done to this. We will give the most general bound between the correctly discriminating probability and that of error in this situation using inequality.(4) and this bound comes to the former results in different special conditions. Consider a quantum system prepared in one of the two states $`|\psi _\pm ^M`$ .We are not told which of the states the system is in, although we do know that it has some probability of being in either. Denote $`P_{IP}=\psi _+\psi _{}^M`$.We represent $`X`$, $`\mathrm{\Gamma }`$ and $`T`$ as $`X=\left(\begin{array}{cc}1& P_{IP}\\ P_{IP}^{}& 1\end{array}\right)`$, $`\mathrm{\Gamma }=diag(\sqrt{P_D^+},\sqrt{P_D^{}})`$, $`T=\left(\begin{array}{cc}0& \sqrt{P_E^+}\\ \sqrt{P_E^{}}& 0\end{array}\right)`$ and inequality.(4) yield $$(1P_D^+P_E^+)(1P_D^{}P_E^{})(P_{IP}\sqrt{P_D^+P_E^{}}\sqrt{P_D^{}P_E^+})(P_{IP}^{}\sqrt{P_D^+P_E^{}}\sqrt{P_D^{}P_E^+}).$$ (8) For $`P_I^+=1P_D^+P_E^+`$, $`P_I^{}=1P_D^{}P_E^{}`$, we can rewrite inequality.(8) as $$P_I^+P_I^{}\left|P_{IP}\sqrt{P_D^+P_E^{}}\sqrt{P_D^{}P_E^+}\right|^2.$$ (9) This inequality is just the most general bound among the discriminating probabilities of correct ($`P_D^+`$, $`P_D^{}`$),error ($`P_E^+`$, $`P_E^{}`$) and inconclusive ($`P_I^+`$, $`P_I^{}`$). In the following we will give some special conditions and simple inequality.(9). 1. We let $`P_I^+=P_I^{}=0`$, which means that we don’t give inconclusive results, then inequality.(9) yield $$P_{IP}=\sqrt{P_D^+(1P_D^{})}+\sqrt{P_D^{}(1P_D^+)}.$$ (10) Eq.(10) give a bound of the maximum correctly discriminating probabilities $`P_D^+`$, $`P_D^{}`$ of the two states $`|\psi _+`$ and $`|\psi _{}`$ given $`M`$ initial copies of each states. We find $`P_{IP}`$ must be real, which means if the inter-produce of two states $`|\psi _+`$ and $`|\psi _{}`$ is not real we can’t execute discrimination without inconclusive results. Furthermore we can suppose $`P_D^+=P_D^{}=P_D`$, and derive $$P_D=\frac{1}{2}(1+\sqrt{1P_{IP}^2}).$$ (11) When $`M=1`$, Eq.(11) just give the well-known Helstrom limit $$P_D=P_H=\frac{1}{2}(1+\sqrt{1\left|\psi _+\psi _{}\right|^2}).$$ (12) Eq.(11) give the absolute maximum probability of discriminating between two states $`|\psi _\pm `$ with given $`M`$ initial copies. The measurement it represents does not give inconclusive results, but will incorrectly identify the states with probability $`1P_D`$. 2. We suppose $`P_{IP}=P_{IP}^{}`$, $`P_D^+=P_D^{}=P_D`$, $`P_E^+=P_E^{}=P_E`$, so $`P_I^+=P_I^{}=P_I`$ and Eq.(9) yield $$\frac{1}{2}(P_{IP}P_I)\sqrt{P_EP_D}\frac{1}{2}(P_{IP}+P_I)\text{ .}$$ (13) Inequality (13) give a general lower bound on the combination of errors and inconclusive result and corresponds to a family of measurements which optimally interpolates between the Helstrom and IDP limits. If $`P_{IP}P_I`$ ,we can get $$\frac{1}{4}(P_{IP}P_I)^2P_EP_D\frac{1}{4}(P_{IP}+P_I)^2\text{ .}$$ (14) When $`M=1`$, the left side of inequality (14) is just the inequality that has been obtained by Chefles and Barnett . If $`P_{IP}<P_I`$, we have $$P_EP_D\frac{1}{4}(P_{IP}+P_I)^2$$ (15) 3. We suppose $`P_E^+=P_E^{}=0`$, and inequality (8) yield $$(1P_D^+)(1P_D^{})\left|P_{IP}\right|^2.$$ (16) So we obtain $$\frac{P_D^++P_D^{}}{2}1\left|\psi _+\psi _{}\right|^M$$ (17) When $`P_D^+=P_D^{}=P_{IDP}`$, we give $`P_{IDP}=1\left|\psi _+\psi _{}\right|^M=1P_{IP}`$, which is a generalization of IDP limit. Denote $`P_S=P_D+P_E`$. It is obvious that $`P_S`$ is the probability with which we measurement the probe system and get $`|P_0`$ after the unitary evolution in Eq.(2). Comparing with , we can say $`P_S`$ is just the probability that two states can be separated by an arbitrary degree with states separating operation. In similar way we can denote $`P_S^{(i)}=P_D^{(i)}+P_E^{(i)}`$ and find the unitary evolution $`\widehat{U}`$ in Eq.(2) just transfer states $`|\mathrm{\Psi }_i_A^M|P_0`$ into states $`\frac{1}{\sqrt{P_S^{(i)}}}\left(\sqrt{\gamma _i}|\phi _i_A+_jt_{ij}|\phi _j_A\right)|P_0`$ with transfer probability $`P_S^{(i)}`$ and realize states separating operate. We can give $`P_D=P_SP_H`$, and $`P_E=P_S(1P_H)`$, where $`P_H`$ is the Helstrom limit for $`M`$ initial states. With these representations, inequality (13) yield $$P_E\frac{1}{2}\left(P_S\sqrt{P_S^2(P_SP_{IDP})^2}\right)\text{ if }P_SP_{IDP}$$ (18) This inequality is just that has been obtained by Chefles and Barnett when $`M=1`$, which give a bound on the error probability $`P_E`$ given a fixed value of the probability $`P_S`$ and it is equivalent to inequality (15). So far we have constructed general discrimination of the states secretly chosen from a certain set $`S`$ given $`M`$ initial copies of each state by a combination of unitary evolution together with measurements. We have derived the matrix inequality, which describe the bound among three different discriminating results: correct, error and inconclusive. For different $`n`$-state,we give a condition which yield such $`n`$-state are linearly independent and find the result of Duan and Guo is just obtained in a special situation of our condition. For two states, we find the most general bound (inequality (9)) among the discriminating probabilities of correctly ($`P_D^+`$, $`P_D^{}`$),error ($`P_E^+`$,$`P_E^{}`$) and inconclusive ($`P_I^+`$,$`P_I^{}`$) and the former works are the results of different applications of our bound in different situation. This work was supported by the National Natural Science Foundation of China.
no-problem/9908/cond-mat9908063.html
ar5iv
text
# 1. Introduction ## 1. Introduction The anomalous transport properties of quasicrystals evince the peculiar electronic properties of these intermetallic alloys, which are reflected in the nature and the density of electronic states in quasiperiodic systems. Tight-binding models on quasiperiodic tilings serve as simple toy models of electrons moving in a quasiperiodic environment. While much is known rigorously for one-dimensional aperiodic Schrödinger operators , quasiperiodic tight-binding models in the physically relevant dimensions two and three have mainly been investigated numerically, compare and references therein for the few exceptions, which, however, yield only rather limited information about the overall structure of the spectrum or the wave functions. At least in two dimensions, numerical results provide ample evidence that typical eigenstates of quasiperiodic tight-binding Hamiltonians are, as in the one-dimensional case, neither extended nor exponentially localized; instead, the probability amplitudes show a multifractal distribution . However, the integrated density of states does not show the hierarchical plateau structure observed for one-dimensional substitution chains ; for pure hopping Hamiltonians it appears to be rather smooth, apart from a few gaps in the spectrum. This makes it possible to employ spectral statistics as further tools to characterize the energy spectrum of quasiperiodic tight-binding models. ## 2. Energy-level statistics The statistical analysis of the spectra of quantum systems has a rich history, and is at the center stage of current interest . For disordered systems, it is by now well known that in the metallic regime the level-spacing distribution $`P_0(s)`$ of neighbouring eigenenergies is well described by $`P_0^{\mathrm{GOE}}(s)`$ of the Gaussian orthogonal ensemble (GOE) of random matrix theory . Here, $`s`$ denotes the energy spacing in units of the mean level spacing, and $`P_0(s)`$ gives the distribution of normalized gaps in the spectrum in the limit of infinite system size. In the insulating regime, the level-spacing distribution follows the Poisson law $`P_0^\mathrm{P}(s)=\mathrm{exp}(s)`$. At the metal-insulator transition, the level-spacing distribution has been shown to follow yet another behavior which is attributed to the existence of a “critical ensemble” . Recently, we have shown that the level statistics of tight-binding models defined on planar quasiperiodic tilings, notably the Ammann-Beenker tiling shown in Fig. 1, is also given by the GOE predictions , contrary to previous statements in the literature . The deviations from GOE behaviour observed in could be understood by noting that the standard periodic approximants of the octagonal tiling are singular patches whose fourfold rotational symmetry is broken only weakly, see Fig. 2. This “almost symmetry” of the patch does not lead to a block structure of the Hamiltonian (which would mean independent spectra for each block), but nevertheless affects the level-spacing distribution. As shown in , the level-spacing distribution $`P_0(s)`$ of generic patches of the tiling, or of the irreducible sectors of patches with exact symmetries, follows the GOE distribution when applying a suitable unfolding procedure to correct for the fluctuations in the density of states. This result was further corroborated by higher correlations such as the Mehta-Dyson statistics $`\mathrm{\Delta }_3`$ and the number variance $`\mathrm{\Sigma }_2`$ . Furthermore, we have shown that the level-spacing distribution remains GOE if one considers small parts of the spectrum without any unfolding, and we demonstrated that the unfolding procedure has to be performed somewhat differently from the procedure in disordered systems . In this paper, we numerically calculate not only the first, but also the second, the third, and the fourth neighbour level-spacing distributions $`P_n(s)`$, $`n=0,\mathrm{},3`$, respectively. We show that the higher level-spacing distributions are also well described by the predictions of random matrix theory. The recently proposed “semi-Poisson ensemble” , supposed to be valid for the multifractal eigenstates of the metal-insulator transition in the three-dimensional Anderson model , does not describe the higher level-spacing distributions of the quasiperiodic models. The same applies to other intermediate statistics obtained recently for an ensemble constructed by an analogy with a one-dimensional gas model of particles with nearest neighbour interactions $`V(x)=\mathrm{log}|x|`$ . This is intermediate between the Poisson and the GOE statistics. which can also be interpreted in terms of Boltzmann distributions of one-dimensional interacting gases with pair potentials $`V(x)=0`$ and $`V(x)=\mathrm{log}|x|`$, but not restricted to nearest neighbours, respectively. Finally, we calculate the participation numbers in these tilings and show that their scaling behavior indicates multifractal behavior. ## 3. Higher level-spacing distributions In addition to the nearest-neighbour level-spacing distribution $`P_0(s)`$ considered previously , we have now computed the distributions $`P_1(s)`$, $`P_2(s)`$, and $`P_3(s)`$ for the simple Hamiltonian $$H=\underset{ij}{\overset{N}{}}t_{ij}|ij|$$ (1) defined on the Ammann-Beenker or octagonal tiling as shown in Fig. 1. Here, each vertex $`i=1,2,\mathrm{},N`$ of the tiling carries a state $`|i`$, and the hopping elements $`t_{ij}=t_{ji}`$ are chosen to be unity between neighbouring vertices, i.e. vertices connected by a bond, and zero otherwise. The spectral statistics are obtained by numerically diagonalizing the Hamiltonian (1) with free boundary conditions for large patches of the tiling, taking into account their symmetries. While the GOE predictions for these distributions are known in principle , it is non-trivial to obtain numerical data for these functions. As we could not find sufficiently accurate results in the literature, we calculated $`P_0^{\mathrm{GOE}}(s)`$, $`P_1^{\mathrm{GOE}}(s)`$, $`P_2^{\mathrm{GOE}}(s)`$, and $`P_3^{\mathrm{GOE}}(s)`$ using the approach of by numerically solving the appropriate differential equations. This was done using Mathematica by performing a small-$`s`$ expansion, followed by a numerical integration to larger values of $`s`$. The small-$`s`$ behaviours of the GOE distributions are $`P_0^{\mathrm{GOE}}(s)=\pi ^2s/6+𝒪(s^3)`$, $`P_1^{\mathrm{GOE}}(s)=\pi ^4s^4/270+𝒪(s^6)`$, $`P_2^{\mathrm{GOE}}(s)=\pi ^8s^8/1764000+𝒪(s^{10})`$, and $`P_3^{\mathrm{GOE}}(s)=\pi ^{12}s^{13}/168781725000+𝒪(s^{15})`$. In our calculations we evaluated the expansions beyond the 30th order. In Fig. 3, we show the results for the data of the largest patch considered in , an octagonal patch as in Fig. 1 but with $`157369`$ vertices, employing the unfolding procedure described in . Apparently, our data agree very well with the GOE distributions. We like to emphasize that the higher level-spacing distributions considered here, and in particular their small-$`s`$ behaviour, are very sensitive to detect deviations from the GOE behaviour. However, we do not find deviations beyond the fluctuations expected for finite systems. Thus, the level-spacing distribution is the same as that of the Anderson model of localization in the metallic regime . This is surprising, because it is well-known that eigenstates in the metallic regime are extended, whereas there is ample evidence that the eigenstates on planar quasiperiodic tilings have a multifractal character. Thus one might speculate that the level statistics should have been similar to the intermediate statistics found at the metal-insulator transition in three dimensions, where eigenstates are multifractal as well. In Fig. 4, we compare our results to the level-spacing distributions of the intermediate semi-Poisson ensemble and the one-dimensional gas ensemble with logarithmic interactions . These latter statistics clearly do not agree with our data. ## 4. Scaling of the participation numbers We now turn our attention to the calculation of the participation numbers $`P(N,E)`$ in order to investigate the properties of the states in the tiling directly. For this purpose, we consider square-shaped patches as in Fig. 1 with up to $`N=7785`$ vertices and compute $$P(N,E)^1=\underset{i=1}{\overset{N}{}}|\psi _i^4(E)|$$ (2) where $`\psi (E)=(\psi _1(E),\mathrm{},\psi _N(E))`$ denotes a normalized eigenstate at energy $`E`$. Thus a state $`\psi (E)`$ that is completely localized at site $`j`$, i.e. $`\psi _i(E)=\delta _{i,j}`$, has $`P(N,E)=1`$, whereas a fully extended state $`\psi _i(E)=1/\sqrt{N}`$ gives $`P(N,E)=N`$. Because for the models under consideration the level-spacing distribution does not appear to depend on the energy , we concentrate on the energy-averaged participation number $`P(N)`$. For multifractal states, we expect to find a power-law behaviour $`P(N)N^\kappa `$ as a function of systems size , between the result for localized ($`\kappa =0`$) and extended ($`\kappa =1`$) states. The result is shown in Fig. 5, which yields a value of $`\kappa 0.87\pm 0.05`$. This value, in agreement with previous results for quasiperiodic tight-binding models , is clearly below the values $`\kappa =1`$ for extended states, albeit rather close, corroborating the multifractal character of the eigenstates in our model, though with rather large characteristic exponents . ## 5. Conclusions We considered the spectral properties of a tight-binding Hamiltonian defined on the Ammann-Beenker tiling. Although this system is not energetically disordered, the level-spacing distributions exhibit level repulsion and are in perfect agreement with the GOE predictions. This is similar as for the energetically disordered Anderson model of localization in the metallic regime , and also for tight-binding models defined on certain random graphs . Nevertheless, the scaling of the energy-averaged participation numbers confirms that the corresponding eigenstates are not extended. Their multifractal character is further corroborated by the system-size dependence of the participation numbers. At first glance, it appears surprising that multifractal states may give rise to GOE level statistics. However, it is known from studies of the metal-insulator transition in more than two dimensions that the critical level statistics becomes similar to the Poissonian behaviour for larger dimensions . Thus, it appears possible that the critical level statistics for the two-dimensional case considered here just coincides with the Gaussian orthogonal random matrix ensemble. This scenario is in perfect agreement with the results presented here. ## Acknowledgements We thank C. A. Tracy for pointing out reference and for suggesting how to compute $`P_n^{\mathrm{GOE}}(s)`$. We also thank U. Gerland and E. B. Bogomolny for stimulating discussions on intermediate statistics.
no-problem/9908/cond-mat9908330.html
ar5iv
text
# Nonequilibrium Growth problems ## I Introduction How to characterize the degree of roughness of a surface as it grows and how the roughness varies in time have evolved into an important topic due to diverse interest in physics, biology, chemistry and in technological applications. One crucial aspect of these nonequilibrium growth processes is the scale invariance of surface fluctuations similar to the scale invariance observed in equilibrium critical point phenomena. Although different kinds of growths may be governed by distinct natural processes, they share a common feature that the surface, crudely speaking, looks similar under any magnification and at various times. This nonequilibrium generalization of scaling involving space and time (called “dynamic scaling”) makes this subject of growth problems important in statistical mechanics. Growth problems are both of near-equilibrium and nonequilibrium varieties, and, therefore, they provide us with a fertile ground to study the differences and the extra features that might emerge in a nonequilibrium situation. Take for example the case of crystal growth. In equilibrium, entropic contributions generally lead to a rough or fluctuating surface, an effect called thermal roughening, but, for crystals, because of the lattice periodicity, a roughening transition from a smooth to rough surface, occurs at some temperature. The nature of the growth of a crystal close to equilibrium expectedly depends on whether the surface is smooth or rough. One can also think of a crystal growth process which is far away from equilibrium by subjecting it to an external drive, for instance by random deposition of particles on the surface. The roughening that occurs in the nonequilibrium case is called kinetic roughening. Is the nature of the surface any different in kinetic roughening? Crystals are definitely not the only example of growth processes; some other examples of such nonequilibrium growths would be the growth of bacterial colonies in a petri dish, sedimentation of colloids in a drop, the formation of clouds in the upper atmosphere, and so on. Note the large variation of length scales of these problems. In many such examples it is difficult if not impossible to think of an equilibrium counterpart. Scale invariance in interface fluctuations implies that fluctuations look statistically the same when viewed at different length scales. A quantitative measure of the height fluctuation (height measured from an arbitrary base) is provided by the correlation function $$C(𝐱,t)=[h(𝐱+𝐱_0,t+t_0)h(𝐱_0,t_0)]^2,$$ (1) where $`x`$ and $`t`$ denote the $`d`$ dimensional coordinate on the substrate and time respectively. The averaging in Eq. (1) is over all $`𝐱_0`$, and, by definition, $`C(𝐱,t)`$ is independent of the choice of the arbitrary base. In simple language, scale invariance then means that when the system is, say, amplified by a scaling $`xbx`$ and $`tb^zt`$, the height fluctuations reveal the same features as the original, upto an overall scale factor. Quantitatively, there exists a generalized scaling $`C(x,t)=b^{2\chi }C(bx,b^zt),`$ (2) where $`b`$ is a scale factor, and $`\chi `$ and $`z`$ are known as the roughening and dynamic exponents which are also universal. As a direct consequence of (2), a scaling form for $`C(x,t)`$ can be obtained by choosing $`b=1/x`$ $`C(x,t)=x^{2\chi }\widehat{C}(t/x^z),`$ (3) a form that also explains the origin of the name “dynamic exponent” for $`z`$. The power law behaviour (as opposed to say exponential decay) of the correlation function implies absence of any scale, neither in space nor in time. All the underlying length scales required to define the problem dropped out of the leading behaviour in Eq. 3 Such a scale invariance is one of the most important features of equilibrium phase transitions and is observed when a parameter, say the temperature, approaches its critical value. However, here there is no special tuning parameter; the scale invariance appears from the interplay of competing processes which in the simplest case can be the surface tension and noise present due to inherent randomness in the growth. There can be, of course, more complex events like a phase transition between surfaces with different roughness but scale invariance (not only of the correlation function but of any physical quantity) is generically preserved in all these surfaces. It is worth emphasizing the enormous simplification that occurs in the scaling description. It is only a very few quantities that define the asymptotic behaviour of the system. Consequently, the idea of studying the universal aspects of growth processes is to classify and characterize the various universality classes as determined by the exponents, e.g., $`\chi `$ and $`z`$, the scaling function and if necessary certain other important universal quantities. At this point, it might be helpful to compare the equilibrium and nonequilibrium cases again. In equilibrium, thanks to thermal energy (or “random kicks” from a heat reservoir), all configurations of a system are accessible and do occur, but no net flow of probability between any two states is expected (called “detailed balance”). Consequently, the knowledge of the states (and the energies) of a system allows one to obtain the thermodynamic free energy by summing over the Boltzmann factors $`\mathrm{exp}(E/k_BT)`$, where $`E`$ is the energy of the state, $`T`$ the temperature and $`k_B`$ is the Boltzmann constant. In a nonequilibrium situation, either or both of the above two conditions may be violated, and the framework of predicting the properties of a system from free energy is not necessarily available. A dynamic formulation is needed. By assigning a time dependent probability for the system to be in a configuration at a particular time, one may study the time evolution of the probability. The equilibrium problem can be viewed from a dynamical point also. This description must give back the Boltzmann distribution in the infinite time steady state limit. This is the Fokker Planck approach. The probabilistic description comes from the ensemble picture where identical copies of the same system exchange energy with the bath independently. An alternative approach which finds easy generalization to the nonequilibrium cases is the Langevin approach where one describes the time evolution of the degrees of freedom, in our example $`h(𝐱,t)`$, taking care of the random exchange of energy by a noise. The dynamics we would consider is dissipative so that the system in absence of any noise would tend to a steady state. However for it to reach the equilibrium Boltzmann distribution in the presence of noise, it is clear that the noise must satisfy certain conditions (Einstein relation) connecting it to the system parameters. The nonequilibrium case does not have any thermodynamic free energy as a guiding light and therefore, there is no requirement to reach the Boltzmann distribution. In the Langevin approach, the noise term can be completely independent. In the equilibrium case the Langevin equation will be determined by the Hamiltonian or the free energy of the system, but for nonequilibrium cases there might be terms which cannot be obtained from Hamiltonians. Since for $`t\mathrm{}`$, the probability distribution for equilibrium cases attains the Boltzmann distribution, the roughness exponent $`\chi `$ is determined even in dynamics by the stationary state while the details of the dynamics is encoded in the dynamic exponent $`z`$. In other words the two exponents $`\chi `$ and $`z`$ are independent quantities. In the nonequilibrium case, there is no compulsion to reach any predetermined stationary state and therefore the surface roughness is related to the growth, i.e., $`\chi `$ and $`z`$ need not be independent. We see below that there is in fact a specific relation connecting these two exponents. The existence of scale invariance and universal exponents implies that as far as the exponents are concerned the theory should be insensitive to the microscopic details or in other words one may integrate out all the small length scale features. The universal exponents come out as an output of this process of coarse graining of say the Langevin equation, followed by a length rescaling that brings the system back to its original form. The new system will however have different values of the parameters and one can study the flow of these parameters in the long length and time scale limit. This is the basic idea behind the renormalization group (RG). In this approach the importance of an interaction or a term is judged not by its numerical value but rather by its relevance. One may start with any physically possible process in dynamics and see how it appears as the length scale or resolution changes. For large length scales, one is left only with the relevant terms that grow with length, and marginal terms that do not change; the irrelevant terms or interactions that decay with length scale automatically drop out from the theory. The exponents are determined at the fixed points of the flows in the parameter space. These fixed points, which remain invariant under renormalization, characterize the macroscopic or asymptotic behavior of the system. Clearly from this view point of renormalization group, one can explain why the microscopic details can be ignored and how the idea of “universality” emerges. All systems whose dynamical behaviour would flow to the same fixed point under RG transformation will have identical scaling behaviour. The various universality classes can then be associated with the various fixed points of RG transformations, and phase transitions or criticality with unstable fixed points or special flows in the parameter space. An RG approach therefore seems rather natural and well suited for studying any scale invariant phenomena in general, growth problems in particular. Quite expectedly, the modern approach to growth problems is based on these views of RG. For a quantitative discussion, we consider two simple equations that, from the historical point of view, played a crucial role in the development of the subject in the last two decades. A simple Langevin equation describing the dynamics of a surface is the Edwards-Wilkinson (EW) equation $`{\displaystyle \frac{h}{t}}=\nu ^2h+\eta (x,t),`$ (4) where $`x`$ represents in general the coordinate on the $`d`$ dimensional substrate. $`\nu `$ is the coefficient of the diffusion term trying to smoothen the surface and $`\eta `$ is the Langevin noise which tries to roughen the surface. One may add a constant current $`c`$ to the right hand side, but by going over to a moving frame of reference ($`hh+ct`$) one recovers Eq. 4. The noise here is chosen to have zero mean and short range correlation as $`\eta (𝐱,t)\eta (𝐱^{},t^{})=2D\delta (𝐱𝐱^{})\delta (tt^{})`$. One of the important assumptions in this equation is that the surface is single-valued and there are no overhangs. One can solve (4) exactly just by taking the Fourier transform and obtain the exponents $`\chi =(2d)/2`$ and $`z=2`$. That the dynamic exponent $`z`$ is 2 follows from the simple fact that the equation involves the first derivative in time but a second derivative in space. The surface is logarithmically rough at $`d=2`$. For $`d>2`$, the fluctuations in the height are bounded and such a surface is more or less flat, better called “asymptotically flat”. From the growth equation one can also derive the stationary probability distribution for the height $`h(𝐱)`$ which takes the form of a Boltzmann factor $`P(h(𝐱))\mathrm{exp}[(\nu /D)(h)^2d^dx]`$ resembling an equilibrium system at a temperature given by $`D=k_BT`$. This is the Einstein relation noise should satisfy to recover equilibrium probability distribution. Conversely, given a hamiltonian of the form $`(h)^2d^dx`$, the equilibrium dynamics will be given by Eq. 4 with $`D`$ determined by the temperature. Nevertheless, if we do not ascribe any thermal meaning to $`D`$, Eq. 4 is good enough to describe a nonequilibrium dynamic process as well. Such a nonequilibrium growth will have many similarities with equilibrium processes, differing only in the origin of the noise, e.g. the expected symmetry $`hh`$ with $`h=0`$ in equilibrium will be preserved in the nonequilibrium case also. The growing surface with a correlation $`C(𝐱,t)=x^{(2d)/2}\widehat{C}(t/|x|^2)`$ will be similar in both cases for $`d<2`$. A genuine nonequilibrium process will involve breaking the up-down symmetry which in equilibrium follws from detailed balance. It should therefore be represented by a term involving even powers of $`h`$. We already saw that a constant current (zeroth power) does not add anything new. Since the origin in space or time or the position of the basal plane should not matter, the first possible term is $`(h)^2`$. By looking at the geometry of a rough surface, it is easy to see that such a term implies a lateral growth that would happen if a deposited particle sticks to the first particle it touches on the surface. One gets the Kardar-Parisi-Zhang (KPZ) equation $`{\displaystyle \frac{h}{t}}=\nu ^2h+{\displaystyle \frac{\lambda }{2}}(h)^2+\eta (𝐱,t).`$ (5) As a consequence of its mapping to the noisy Burger’s equation, to the statistical mechanics of directed polymer in a random medium and other equilibrium and nonequilibrium systems, the KPZ equation has become a model of quite widespread interest in statistical mechanics. Though we focus on growth problems in this paper, the KPZ equation is also applicable in erosion processes. Taking a cue from the development in understanding the growth phenomena through the KPZ equation, a vast class of simulational and analytical models have evolved to explain different experimentally observed growth processes. Diverse technical tools ranging from simulations with various dynamical rules to different versions of renormalization group techniques, mode coupling theory, transfer matrix techniques, scaling arguments have been employed to understand kinetic roughening. In this review we attempt to provide an overview of this phenomenon of roughening of a growing surface. It is almost beyond the scope of this review, to describe in detail various models and their experimental relevance. Rather we focus our attention on a few examples which may broadly represent a few different routes along which research has continued. The plan of this article is as follows. In the next section we focus on the KPZ equation and its renormalization group description. We also point out the connection of the KPZ equation to some other problems of physics. In section III a more generalized growth mechanism involving nonlocal interactions is presented. Section IV is devoted to the progress in understanding the roughening and super roughening transitions which appear in a very distinct class of models involving lattice pinning potential. ## II KPZ equation and more Let us first look at the origin of the various terms in Eq. 4 and 5. In both the equations, the noise term represents random deposition, the fluctuation around the steady value. As already mentioned a steady current can be removed from the equation by going over to a moving frame. The term involving second derivative of $`h`$ can represent either of two processes. It could be a surface tension controlled diffusion process, in which a particle comes to the surface and then does a random walk on the surface to settle at the minimum height position thereby smoothening the surface. An alternative interpretation would be that there is desorption from the surface and the process is proportional to the chemical potential gradient. The chemical potential of the particles on the surface cannot depend on $`h`$ or gradient of $`h`$ because of its independence of the arbitrary base or its tilt. The chemical potential then is related to the second derivative of $`h`$. This also has a geometric meaning that $`^2h`$ is related to the local curvature. The larger the curvature, the higher is the chance to desorb because of a lesser number of neighbours. In the KPZ equation, the nonlinear term represents lateral growth. The diffusion-like term can then be thought of either (a) as an alternative that a particle coming to the surface instead of sticking to the first particle it touches, deposits on the surface and then diffuses, or (b) as a random deposition process with desorption. In either case, the noise term tends to roughen the surface, the diffusion term, of whatever origin, smoothens it while the nonlinear term leads to a laterally growing surface. Even if the smoothening linear term is not present, renormalization group or the scaling argument indicates that such a term is generated on a large length scale. The KPZ equation has a special symmetry not present in the Edwards-Wilkinson case. This is the tilt symmetry (often called Galilean invariance - a misnomer, though, in this context). If we tilt the surface by a small angle, then with a reparametrization $`h^{}=h+ϵ𝐱`$ and $`𝐱=𝐱+\lambda ϵt^{}`$ and $`t=t^{}`$, the equation remains invariant for small $`ϵ`$. This transformation depends only on $`\lambda `$ the coefficient of the nonlinear term and fails for $`\lambda =0`$. Since this tilt symmetry is to be maintained no matter at what lengthscale we look at, $`\lambda `$ must be a renormalization group invariant. Let us now perform a length rescaling analysis. Under a change of scale as $`xbx`$, $`tb^zt`$ and $`hb^\chi h`$, KPZ equation transforms as $`b^{\chi z}{\displaystyle \frac{h}{t}}=\nu b^{\chi 2}^2h+{\displaystyle \frac{\lambda }{2}}b^{2\chi 2}(h)^2+b^{d/2z/2}\eta ,`$ (6) where the noise correlation has been used to obtain the scaling of the noise term. Therefore under this scale transformation different parameters scale as $`\nu b^{z2}\nu `$, $`Db^{zd2\chi }D`$ and $`\lambda b^{\chi +z2}\lambda `$. For $`\lambda =0`$, the equation remains invariant provided $`z=2`$ and $`\chi =(2d)/2`$. These are just the exponents one expects from the EW model. (Such surfaces with anisotropic scaling in different directions like $`x`$ and $`h`$ are called self-affine. ) Though we cannot predict the exponents from Eq. 6 when $`\lambda 0`$, it does tell us that a small nonlinearity added to the EW equation scales with a scaling dimension $`\chi +z2`$. This term is always relevant in one dimension, because it scales like $`b^{1/2}`$. This type of scaling argument also shows that no other integral powers of derivatives of $`h`$ need be considered in Eq. 5 as they are all irrelevant, except $`(h/x)^3`$ at $`d=1`$, which, however, detailed analysis shows to be marginally irrelevant. Based on this analysis, we reach an important conclusion that the nonequilibrium behaviour in one dimension, and in fact for any dimensions below two, would be distinctly different from the equilibrium behaviour. For dimensions greater than two, Edwards Wilkinson or equilibrium surfaces, as already mentioned, are asymptotically flat with $`\chi =0,z=2`$, and so, a small nonlinearity is irrelevant because it will decay with $`b`$. In other words, the growth in higher dimensions for small $`\lambda `$ would be very similar to equilibrium problems because the EW model is stable with respect to a small perturbation with nonlinearity. The simple scaling argument does not tell us if the nature of the surface changes for large $`\lambda `$ for $`d>2`$, but an RG analysis shows that it does change. That the nonequilibrium growth is always different in low dimensions and in higher dimensions (greater than two), and that there will be a dynamic phase transition from an equilibrium-like to a genuine nonequilibrium behaviour, explains the source of excitement in this minimal KPZ equation, in the last two decades. If the nonlinear parameter $`\lambda `$ is to remain an invariant, i.e. independent of $`b`$ in Eq. (6), then $`\chi +z=2`$, a relation which need not be satisfied by the equilibrium growth. It is this relation connecting the two exponents of the scaling function of Eq. 2 that distinguishes nonequilibrium growth from equilibrium, the former requiring one less exponent than the latter one. We wonder if such an exponent relation is generally true for all nonequilibrium systems. Though we are far away from a complete understanding of all the nuances and details of the KPZ equation, the renormalization group analysis has been very successful in identifying different phases, nature of phase transitions, and, in certain cases, relevant exponents. In brief, the various results obtained from renormalization group analysis are as follows. In one dimension and for $`d<2`$, even a small nonlinearity, as already mentioned, being relevant in the RG sense, leads to new values of roughening and dynamic exponents, and is characterized by a RG fixed point. Beyond $`d=2`$, there is a phase transition demarcating two different types of surfaces. A small nonlinearity is irrelevant around EW model and the surface is almost flat with $`\chi =0`$ and $`z=2`$. A strong nonlinear growth, however, drives the system to a different phase with rougher surface where $`\chi 0`$. Several aspects of this phase transition can be studied from RG but the strong $`\lambda `$ regime is still out of reach, because of the absence of any RG fixed point. The KPZ equation in $`d=1`$ has distinct nonequilibrium behaviour, and the scaling behaviour is the same no matter how small or large $`\lambda `$ is. More peculiar is the existence of a stationary probability distribution of the height in one dimension which is the same as for the linear EW model. This is not just an accident but a consequence of certain subtle relations valid only in one dimension. We do not go into those issues here. The same stationary distribution implies that the nonlinearity does not affect the stationary state solution, and $`\chi =1/2`$. The two models however differ in the dynamic exponent which, in the case of KPZ growth, has to satisfy $`\chi +z=2`$. This leads to exact answer $`z=3/2`$. Its significance can be grasped if we compare various known cases. For ballistic motion, distance goes linearly with time so that the dynamic exponent is $`z=1`$ while for diffusive motion or in quantum mechanics (e.g. a nonrelativistic free quantum particle), $`z=2`$ as also the case for EW. Here is an example where the nonequilibrium nature of the problem leads to a completely new exponent connecting the scaling of space and that of time. ### A Dynamic renormalization group analysis A dynamic renormalization group analysis is a more general approach applicable for dynamics which e.g. may be governed by the Langevin equation for the appropriate dynamical variable. For our problem it is easier to work in Fourier coordinates $`𝐪`$, and $`\omega `$ conjugate to space and time. Long distance, long time implies $`𝐪0`$ and $`\omega 0`$, and $`q`$ can be taken as the inverse wavelength at which the height variable is probed. The magnitude of wave vector $`𝐪`$ varies from $`0`$ to $`\mathrm{\Lambda }`$ where the upper cutoff is determined by the underlying microscopic length scale like lattice spacing or size of particles etc. In the Fourier space, different Fourier modes in the linear EW model gets decoupled so that each $`h(𝐪,\omega )`$ for each $`(𝐪,\omega )`$ behaves independently. It is this decoupling that allows the simple rescaling analysis of Eq. 6 or dimensional analysis to give the correct exponents. For the KPZ equation the nonlinear term couples heights of various wavelengths and therefore any attempt to integrate out the large $`(𝐪,\omega )`$ modes will affect $`h`$ with low values of $`(𝐪,\omega )`$. This mixing is taken into account in the RG analysis which is implemented in a perturbative way. One thinks of the noise and the nonlinear term as disturbances affecting the EW-like surface. If we know the response of such a surface to a localized disturbance we may recover the full response by summing over the disturbances at all the points and times. However this disturbance from the nonlinear term itself depends on the height, requiring an iterative approach that generates successively a series of terms. By averaging over the noise, one then can compute any physical quantity. At this stage only degrees of freedom with $`𝐪`$ in a small shell $`e^l\mathrm{\Lambda }<q<\mathrm{\Lambda }`$ is integrated out. In real space this corresponds to integrating out the small scale fluctuation. The contribution from this integration over the shell is absorbed by redefining the various parameters $`\nu `$, $`\lambda `$ and $`D`$. These are the coupling constants for a similar equation as (5) but with a smaller cutoff $`\mathrm{\Lambda }e^l`$. A subsequent rescaling then restores the original cutoff to $`\mathrm{\Lambda }`$. Following this procedure, the flow equations for different parameters $`\nu `$, $`D`$, and $`\lambda `$ can be obtained. Using the exponent identity predicted from the Galilean invariance and the renormalization group invariance of $`\nu `$, the flow equations for all the parameters can be combined into a single flow equation for $`\overline{\lambda }^2=\lambda ^2D/\nu ^3`$ (with $`\mathrm{\Lambda }=1`$). This is the only dimensionless parameter that can be constructed from $`\lambda `$, $`\nu `$, $`D`$, and $`\mathrm{\Lambda }`$, and it is always easier to work with dimensionless quantities. Its recursion relation is $`{\displaystyle \frac{d\overline{\lambda }}{dl}}={\displaystyle \frac{2d}{2}}\overline{\lambda }+K_d{\displaystyle \frac{2d3}{4d}}\overline{\lambda }^3,`$ (7) where $`K_d`$ is the surface area of a $`d`$-dimensional sphere divided by $`(2\pi )^d`$. The invariance of $`\nu `$ under RG transformation implies $`z=2K_d\overline{\lambda }^2\frac{2d}{4d},`$ and the Galilean invariance provides the value of $`\chi =2z`$ once the value of $`z`$ is known. To be noted here is that the dynamic exponent is different from 2 by a term that depends on $`\lambda `$ coming from the renormalization effects. A few very important features are apparent from (7). From the fixed point requirement $`d\overline{\lambda }/dl=0`$, we find that at $`d=1`$, there is a stable fixed point $`\overline{\lambda }^2=2/K_1`$. At this fixed point $`z=3/2`$ and $`\chi =1/2`$ supporting the results predicted from the symmetry analysis. At $`d=2`$, the coupling is marginally relevant indicating a strong coupling phase not accessible in a perturbation scheme. At $`d>2`$, the flow equation indicates two distinct regimes, namely a weak coupling regime where $`\overline{\lambda }`$ asymptotically vanishes leading to a flat EW phase with $`\chi =0,z=2`$, and a strong coupling rough phase, the fixed point of which cannot be reached by perturbation analysis. Owing to this limitation of the renormalization group analysis based on the perturbation expansion, the scaling exponents in this strong coupling phase cannot be determined by this RG scheme. Different numerical methods yield $`z=1.6`$ at $`d=2`$. The phase transition governed by the unstable fixed point of $`\overline{\lambda }`$ is well under control with $`z=2`$ for all $`d>2`$. To explore the strong coupling phase, recently techniques like self-consistent mode coupling approach, functional renormalization group etc have been employed, but even a basic question whether there is an upper critical dimension at which $`z`$ will again become 2 remains controversial. ### B Relation with other systems The relation of the KPZ equation with other quite unrelated topics in equilibrium and nonequilibrium statistical mechanics is impressive. Here we provide a very brief account of these systems. Noisy Burgers equation: By defining a new variable $`𝐯=h`$, we obtain an equation $`{\displaystyle \frac{𝐯}{t}}=D^2𝐯+\lambda 𝐯v+𝐟(𝐱,t),`$ (8) where the noise term $`f=\eta `$. The above equation represents the noisy Burgers equation for vortex free ( $`\times 𝐯=0`$) fluid flow with a random force. This equation is very important in studies of turbulence. The tilt invariance of the KPZ equation turns out to be the conventional Galilean invariance for the Burgers equation (for $`\lambda =1`$), and that’s how the name stayed on. Directed polymer in a random medium: A directed polymer, very frequently encountered in different problems in statistical mechanics, is a string like object which has a preferred longitudinal direction along which it is oriented, with fluctuations in the transverse direction. The flux lines in type II or high T<sub>c</sub> superconductors are examples of such directed polymers in 3 dimensions, while the steps on a vicinal or miscut crystal surface or the domain walls in a uniaxial two dimensional system are examples in two dimensions. The formal mathematical mapping to such objects follows from a simple (Cole-Hopf) transformation of the KPZ equation using $`W(𝐱,t)=\mathrm{exp}[\frac{\lambda }{2\nu }h].`$ The Cole-Hopf transformation linearizes the nonlinear KPZ equation and the resulting linear diffusion equation (or imaginary time Schroedinger equation) is identical to that satisfied by the partition function of a directed polymer in a random potential. For such random problems, one is generally interested in the averages of thermodynamic quantities like the free energy and we see that the noise averaged height $`h(𝐱,t)`$ gives the average free energy of a directed polymer of length $`t`$ with one end at origin and the other end at $`𝐱`$. This is a unique example of a system where the effect of such quenched averaging of free energy can be studied without invoking any tricks (like the replica method). This has led to many important results and enriched our understanding of equilibrium statistical mechanics. Recently, this formulation has been extended to study details of the properties of the random system near the phase transition point and overlaps in lower dimensions. It turns out that one needs an infinite number of exponents to describe the statistical behaviour of the configurations of the polymer in the random medium. We do not go into this issue as this is beyond the scope of this article. An interesting connection between the $`1+1`$ dimensional KPZ equation and the equilibrium statistical mechanics of a two dimensional smectic -A liquid crystals has been recently established by Golubovich and Wang. This relationship further provides exact approach to study the anomalous elasticity of smectic-A liquid crystals. Apart from these, there are a number of other relations between KPZ equation and kinetics of annihilation processes with driven diffusion, the sine-Gordon chain, the driven diffusion equation and so on. ### C Beyond KPZ #### 1 Conservation condition The situation encountered in Molecular-beam epitaxy (MBE) for growth of thin films is quite different than the mechanism prescribed by the KPZ equation. In MBE the surface diffusion takes place according to the chemical potential gradient on the surface, respecting the conservation of particles. If the particle concentration does not vary during growth, then a mass conservation leads to a volume conservation and the film thickness is governed by an underlying continuity equation $`{\displaystyle \frac{h}{t}}+𝐣=\eta ,`$ (9) where $`j`$ is the surface diffusion current which states that the change of height at one point is due to flow into or out from that point. The current is then determined by the gradient of the chemical potential, and since the chemical potential has already been argued to be proportional to the curvature $`^2h`$, the growth equation thus becomes a simple linear equation involving $`^4h`$ which, like the EW model, is exactly solvable. Taking into account the effect of nonlinearity the full equation can be written as $`{\displaystyle \frac{h}{t}}=^2[\nu ^2h+{\displaystyle \frac{\lambda }{2}}(h)^2]+\eta (x,t),`$ (10) where the noise correlation is $`\eta (𝐱,t)\eta (𝐱^{},t^{})=2D^2\delta (𝐱𝐱^{})\delta (tt^{})`$, if the noise also maintains conservation (if it originates from the stochasticity of diffusion) or would be the same white noise as in the KPZ equation if the noise is from random deposition. It goes without saying that the exponents are different from the EW model even for the linear theory. The invariance of $`\lambda `$ in this case leads to a different relation between $`\chi `$ and $`z`$. At the dimension of physical interest $`d=2`$, this growth equation leads to an enhanced roughness than the KPZ case and may explain the results of experiments of high temperature MBE. #### 2 Quenched noise A different type of generalization of the KPZ equation was to explore the motion of domain walls or interfaces in a random medium. In this case, the noise is not explicitly dependent on time but on the spatial position and the height variable. Such a noise has been called quenched noise because the noise is predetermined and the interface or the surface moves in this random system. The simple features of the KPZ equation and the EW model are lost. Functional renormalization group analysis and numerical studies attempted to clarify the question of universality classes and details of dynamics in such cases. The important concept that emerged in this context is the depinning transition so that the surface remains pinned by the randomness until the drive exceeds a certain critical value. Interface depinning is an example of a nonequilibrium phase transition. The velocity of the surface near this depinning transition also has critical like behaviour with long range correlations. Below the threshold, the dynamics is sluggish, while just above the threshold , the velocity is in general not proportional to the drive but obeys a power law with a universal exponent. For a very strong drive (or large velocity of the interface) the moving surface encounters each site only once, and therefore the noise is effectively like a space-time dependent noise rather than the quenched one. The nature of the surface would then be like KPZ. #### 3 Coloured noise In the previous section we discussed the KPZ equation with white noise. If the noise is coloured in the sense that there is correlation in space or time or both, the universal behaviour, the phase transitions and the properties are different but still can be studied by the same RG technique. Several aspects of the problem especially the role of noise correlation have been explored. All of the above seem to suggest that if there is no conservation law, then the KPZ equation is the equation to describe any nonlinear or nonequilibrium growth process and all phenomena can be put in one of the known universality classes. However, experimentally KPZ exponents seem to elude us so far. Since results are known exactly in one dimension, special one dimensional experiments were conducted like paper burning, interface motion in paper, colloid suspension etc, but KPZ exponents have not been seen. In the colloid experiment the surface formed by the depositing colloids on the contact line $`(d=1)`$ between the colloid latex film and a glass slide was measured from video images. This method yields $`\chi =.71`$ but cannot determine the dynamic exponent. A recent analysis of tropical cumulus cloud in the upper atmosphere, from satellite and space shuttle data from 0.1 to 1000 Km, seems to agree with the KPZ results in $`d=2`$. ## III Kinetic roughening with nonlocality In spite of a tremendous conceptual and quantitative success of KPZ equation in describing the nonequilibrium growth mechanism, the agreement with experimentally observed exponents is rather unsatisfactory. One wonders whether there is any relevant perturbation that drives the systems away from the KPZ strong coupling perturbation. One goal of this section is to point out that indeed there can be longrange interactions that may give rise to non-KPZ fixed points. Many recently studied systems involving proteins, colloids, or latex particles the medium induced interactions are found to play an important role. This nonlocal interaction can be introduced by making a modification of the nonlinear term in the KPZ equation. Taking the gradient term as the measure of the local density of deposited particles, the long range effect is incorporated by coupling these gradients at two different points. The resulting growth equation is a KPZ-like equation with the nonlinear term modified as $`\frac{1}{2}𝑑𝐫^{}𝒱(r^{})`$ $`h(𝐫+𝐫^{},t)h(𝐫𝐫^{},t)`$. For generality we take $`𝒱(r^{})`$ to have both short and long range parts with a specific form in Fourier space as $`𝒱(k)=\lambda _0+\lambda _\rho k^\rho `$ such that in the limit $`\lambda _\rho 0`$ KPZ results are retrieved. The aim is to observe whether the macroscopic properties are governed by only $`\lambda _0`$ and hence KPZ like or the behavior is completely different from KPZ due to the relevance of $`\lambda _\rho `$ around the KPZ fixed points. A scaling analysis as done in Eq. (6) clearly indicates different scaling regimes and the relevance of $`\lambda _0`$ and $`\lambda _\rho `$ for $`d<2`$ at the EW fixed point. For any $`\lambda _\rho `$($`0`$) with $`\rho >0`$, the local KPZ theory (i.e. $`\lambda _\rho =0`$ and $`\chi +z=2`$) is unstable under renormalization and a non-KPZ behavior is expected. For $`2<d<2+2\rho `$, only $`\lambda _\rho `$ is relevant at the EW fixed point. The exponents of the non-KPZ phases can be obtained by performing a dynamic renormalization group calculation. By identifying the phases with the stable fixed points, we then see the emergence of a new fixed point where the long range features dominate $`(\chi +z=2+\rho )`$. Most importantly, at $`d=2`$, the marginal relevance of $`\overline{\lambda }`$ is lost and there is a stable fixed point (LR) for $`\rho >.0194`$. On the experimental side, there are experiments on colloids with $`\chi =0.71`$ which is the value also obtained from paper burning exponents. For colloids, hydrodynamic interactions are important. Similar longrange interactions could also play a role in paper burning experiment due to the microstructure of the paper. With this $`\chi `$ our exponents suggest $`\rho =0.12`$ at $`d=1`$ at the long range fixed point. Further experiments on deposition of latex particles or proteins yielding the roughness of growing surface have not been performed. Probably such experiments may reveal more insights on this growth mechanism. More recently, the effect of coloured noise in presence of nonlocality has been studied and the nature of the phases and the various phase transitions clarified. A conserved version of the nonlocal equation has also been considered and it shows rich behaviour. ## IV Roughening transition in nonequilibrium It is interesting to study the impact of equilibrium phase transitions on the nonequilibrium growth of a surface. This is the situation observed experimentally in growth of solid $`{}_{}{}^{4}He`$ in contact with the superfluid phase. There is an equilibrium roughening transition at $`T_R=1.28K`$. For $`T>T_R`$ the growth velocity is linear in the driving force $`F`$ (chemical potential difference), but for $`T<T_R`$ the velocity is exponentially small in the inverse of the driving force. For infinitesimal drive , the mobility which is the ratio of the growth velocity and $`F`$ vanishes with a jump from a finite value at the transition. With a finite force the transition is blurred and the flat phase below $`T_R`$ in equilibrium becomes rougher over large length scale. The equilibrium roughening transition is an effect of discrete translational symmetry of the lattice. The equilibrium dynamics in this case is essentially governed by the Langevin equation $`{\displaystyle \frac{h}{t}}=K^2h(𝐫,t)V\mathrm{sin}[{\displaystyle \frac{2\pi }{a}}h(r,t)]+\zeta (r,t),`$ (11) where the sin term favours a periodic structure of spaoing $`a`$. Extensive investigations have been done on this equilibrium model. At low temperature this periodic potential is relevant and it ensures that minimum energy configuration is achieved when $`\varphi `$ is an integer multiple of lattice periodicity. In this phase the surface is smooth and the roughness is independent of length. In the high temperature phase the equilibrium surface is thermally rough and the roughness is logarithmic $`C(L,\tau )\mathrm{ln}[Lf(\tau /L^z)].`$ (12) The critical point is rather complicated and goes by the name of Kosterlitz-Thouless transition, first discussed in the context of defect mediated transitions in two dimensional XY magnets. For a nonequilibrium crystal growth problem, one needs to introduce the KPZ nonlinear term in (11). There is no longer any roughening transition. The fact that away from equilibrium the roughening transition is blurred is manifested by the domination of the nonlinear term and the suppression of the pinning potential in the asymptotic regime. A very nontrivial situation arises when the surface contains quenched disorder which shifts the position of the minima of the pinning potential in an arbitrary random fashion. In this case there is a new phase transition which is drastically different from the equilibrium roughening transition. This transition is called super roughening. Above the transition temperature i.e. for $`T>T_{sr}`$, the surface is logarithmically rough as it is in the high temperature phase of the pure problem. However in the low temperature phase i.e. for $`T<T_{sr}`$, the surface is no longer flat and is even rougher than the high temperature phase. Recent numerical treatments suggest that the surface roughness behaves as $`(\mathrm{ln}L)^2`$. In the nonequilibrium situation the linear response mobility vanishes continuously at the transition temperature unlike the jump discontinuity in the pure case. A general treatment with a correlated disorder elucidates the connection between the roughening and super roughening transition and one observes that the roughening turns into a super roughening transition if the disorder correlation decays sufficiently fast. Away from equilibrium, the super roughening transition is essentially dominated by the KPZ nonlinearity and instead of the logarithmic roughness, an asymptotic power law behavior of the roughness is found over all temperature range. In a similar situation in the nonequilibrium case, one needs to study the role of the KPZ nonlinearity with long range disorder correlation. A functional renormalization scheme with an arbitrary form of the disorder correlation turns out to be useful, though a detailed solution is not available. It is found that the flow of the KPZ nonlinearity under renormalization, with power law form of the disorder correlation, is such that it decays with length. This implies that nonequilibrium feature does not set in over a certain length scale. Over this scale one would then expect usual roughening transition. However there is a generation of a driving force due to the nonlinearity, and the growth of this force with length scale would invalidate use of perturbative analysis. For large length scales, one expects a KPZ type power law roughness of the surface. Nevertheless, the initial decay of the nonlinearity with the length scale due to the long range correlation of the disorder is an interesting conclusion that seems to be experimentally detectable. ## V Remarks In this brief overview, we attempted to focus on the difference between equilibrium and nonequilibrium growth problems with an emphasis on the scaling behaviour and renormalization group approach. Many details with references to pre-1995 papers can be found in Ref. , which should be consulted for more detailed analysis. Though the success story of the KPZ equation is rather impressive, there are still many unresolved, controversial issues. In fact for higher dimensions, the behaviour is not known with as much confidence as for lower dimensions. Developments in this direction are awaited. Note added in Proof (1) The growth mechanism of metal-organic films deposited by the Langmuir-Blodgett technique has been studied in Ref. 23 by X-ray scattering and atomic force microscopy. The results have been interpreted by a combination of 1-dimensional EW equation (Eq. 4 ) and 2-dimensional linear conserved equation (Eq. 10) with conserved noise. (2) For effects of nonlocality in equilibrium critical dynamics, see Ref. 24.
no-problem/9908/math9908079.html
ar5iv
text
# On the structure of varieties with degenerate Gauss mappings ## I. Joins Form the join of $`k`$ varieties $`Y_1,\mathrm{},Y_kV`$, $$X=S(Y_1,\mathrm{},Y_k)=\overline{_{y_jY_j}_{y_1,\mathrm{},y_k}}$$ where $`_{y_1,\mathrm{},y_k}`$ denotes the projective space spanned by $`y_1,\mathrm{},y_k`$ (generically a $`^{k1}`$). Note that $`\text{dim}X\mathrm{\Sigma }_j\text{dim}Y_j+(p1)`$ with equality expected. Joins have degenerate Gauss maps with at least $`(k1)`$-dimensional fibers because Terracini’s lemma (see \[Z, II.1.10\]) implies that the tangent space to $`S(Y_1,\mathrm{},Y_k)`$ is constant along each $`_{y_1,\mathrm{},y_k}^{k1}`$. Two special cases of this construction: 1. Let $`L`$ be a linear space. Then $`S(Y,L)`$ is a cone over $`Y`$ with vertex $`L`$. 2. $`Y_j=Y`$ for all $`j`$. Then $`X`$ is the union of the secant $`^{k1}`$’s to $`Y`$. Joins are built out of cones in the sense that one can use e.g., the family of cones over $`Y_2`$ with vertices the points of $`Y_1`$ to sweep out $`X`$. ## II. Varieties built from tangent lines ## IIA. Tangential varieties Let $`YV`$ be a variety and let $`\tau (Y)V`$ denote the union of tangent stars to $`Y`$. (If $`Y`$ is smooth, $`\tau (Y)`$ is the union of embedded tangent lines to $`Y`$.) $`\tau (Y)`$ has a degenerate Gauss map with at least one dimensional fibers. (see \[L\] for definitions). One can also take higher osculating varieties of $`Y`$ which will also have degenerate Gauss mappings. Examples IIB and IIC below generalize $`\tau (Y)`$. ## IIB. Hyperbands Let $`YV`$ be a smooth variety of dimension $`m`$ and fix $`k=Nm1`$. For each $`yY`$, let $`L_y𝔾(m+k,V)`$ be such that $`\stackrel{~}{T}_yYL_y`$ and let $`X=_{yY}L_y`$. Then $`\text{dim}XN1`$ (with equality occurring generically) and $`X`$ will have degenerate Gauss map with at least one-dimensional fibers. Such a variety $`X`$ is called a hyperband (see \[AG, p. 255\]). The hyperbands with fibers of dimension greater than one are not built by families of cones and developable varieties. So, they are not covered by the Griffiths-Harris structure theorem. One could seek to generalize tangential varieties in a different way, namely by taking a subspace of the tangent lines through each point of $`Y`$. If $`xV`$ and $`vT_xV`$, we let $`_{x,v}^1`$ denote the line passing through $`x`$ with tangent space spanned by $`v`$. Let $`\mathrm{\Delta }TY`$ be a distribution. One could consider the variety $`X=_{yY,v\mathrm{\Delta }_y}_{y,v}^1`$ consisting of the union of tangent lines tangent to $`\mathrm{\Delta }`$. In general $`X`$ will not have a degenerate Gauss map, but it will in some special cases. The case where $`YZ`$ and one takes $`X=_{yY}\stackrel{~}{T}_yZ`$ is one special case. Here is another construction: ## IIC. Unions of conjugate spaces Let $`II=II_{Y,y}S^2T_y^{}YN_yY`$ denote the projective second fundamental form of $`Y`$ at $`y`$ (see \[AG\], \[GH\] or \[L\] for a definition). Let $`Y^{n1}^{n+1}`$ be a variety such that at general points there exist $`n1`$ simultaneous eigen-directions for the quadrics in its second fundamental form. This condition holds for generic varieties of codimension two. (To make the notion of eigen-direction precise, choose a nondegenerate quadric in $`II`$ to identify $`T`$ with $`T^{}`$ and consider the quadrics as endomorphisms of $`T`$. The result is independent of the choices.) Let $`X^n^{n+1}`$ be the union of one of these families of embedded tangent lines. The directions indicated above are called conjugate directions on $`Y^{n1}`$. In higher dimensions it is still possible to have a conjugate direction or conjugate space, but in this case $`Y`$ must satisfy a certain exterior differential system. As is shown in \[AG, p. 85\] local solutions to this system exist and depend on $`n(n1)`$ arbitrary functions of two variables. In this case $`X`$ is the union of the tangential varieties of the integral curves for the distribution defined by the conjugate directions to $`Y`$. ## III. Varieties with $`f=1`$ ## IIIA. Generic varieties with $`f=1`$ We say a variety $`XV`$ with $`f=1`$ is generic among varieties with $`f=1`$ if $`Z_F`$ consists of $`n1`$ distinct points and the variety each point sweeps out is $`(n1)`$-dimensional. The following theorem follows from results in \[AG\]: ###### Theorem The varieties $`X^n^{n+a}`$ generic among varieties with $`f=1`$ are the union of conjugate lines to some variety $`Y^{n1}`$, with a finite number of lines tangent to a general point of $`Y`$. ## IIIB. Classification of $`X^3^4`$ with $`f=1`$ Here $`F`$ is a $`^1`$ and the focus $`Z_F`$ is of degree two. There are two classes: Class 1: $`Z_F`$ consists of two distinct points, $`z_1,z_2`$. 1a. (Generic case) Each $`z_j`$ traces out a surface $`S_j`$. Here $`X`$ is the dual variety of a $`II`$-generic surface in $`^4`$ (its Gauss image). Locally $`X`$ may be described as the union of a family of lines tangent to conjugate directions on either surface (one must take the family that corresponds to conjugate directions on the other surface). It may be the case that globally $`S_1=S_2`$ and there is a unique construction. 1b. $`z_1`$ traces out a surface $`S`$ and $`z_2`$ traces out a curve $`C`$. Here $`X`$ may be described as the union of a family of conjugate lines to $`S`$, where the conjugate lines intersect along a curve. 1c. Both $`z_j`$’s trace out curves, $`C_j`$. In this case $`X=S(C_1,C_2)`$. Class 2: $`Z_F`$ is a single point $`z`$ of multiplicity two. 2a. $`z`$ traces out a surface $`S`$. In this case $`S`$ will have a family of asymptotic lines and $`X`$ is the union of the asymptotic lines to $`S`$. 2b. $`z`$ traces out a curve $`C`$. An example of $`X`$ in this case is the union of a family of planes that are tangent to $`C`$. We conjecture that this is the only example. 2c. $`z`$ is fixed, then $`X`$ is a cone over $`z`$. Acknowledgement. We are grateful to J. Piontkowski for pointing out that our interpretation of the Griffiths-Harris announcement in an earlier version of this paper was too narrow, and showing us how some of the examples above fit into the Griffiths-Harris perspective.
no-problem/9908/astro-ph9908101.html
ar5iv
text
# Flares on the Black Holes ## 1 Introduction We show that a Kerr black hole can flare quasi-periodically, if it is connected to the accretion disk by a magnetic field. The averaged luminosity in the flares is $`B^2r_+^2c`$, where $`B`$ is the magnetic field at the event horizon, $`r_+`$ is the event horizon radius, $`c`$ is the speed of light. The mechanism considered here is different from the Blandford-Znajek process - the magnetic field connects the hole to the accretion disk, instead of to infinity. In §2 we give the astrophysical formulation of the problem and an order-of-magnitude solution. In §3 we idealize the problem and solve it rigorously. ## 2 Astrophysics of Black Hole Flares Fig. 1. shows a Kerr black hole, its accretion disk, and a magnetic spot in the disk. The hole will twist the magnetic field lines, that is a toroidal magnetic field will be generated. To calculate the toroidal field, we treat the hole as a rotating conducting ball, radius $`r_+`$, conductivity $`\sigma c/r_+`$ (Znajek 1978). We assume for simplicity that the ball is rotating at about the speed of light, corresponding to a nearly maximally rotating black hole. In the ball rest frame, the poloidal electric field is $`EB`$, where $`B`$ is the magnetic field. The current is $`Ir_+^2\sigma Ecr_+B`$. This current creates a toroidal magnetic field, $`B_tI/(cr_+)B`$. The twisted magnetic field will try to unwind itself. According to the Kruskal-Shafranov criterion (Kadomtsev 1966), this leads to the screw instability when the toroidal field becomes so strong that the magnetic field line turns around itself about once. The instability condition is $`B_t/r_+B/L`$, or $`Lr_+`$, where L is the distance from the event horizon to the magnetic spot measured along the field line. If the stationary magnetic configuration is screw-unstable, the black hole magnetosphere flares and releases part of the magnetic energy. Then the hole twists the magnetic field again, and the magnetosphere becomes screw-unstable. The flares repeat quasi-periodically, with the period $`r_+/c`$. ## 3 Idealized problem We straighten the field lines connecting the hole and the magnetic spot in the disk. We neglect the motion of the spot as compared to the hole rotation. This transforms Fig.1 into Fig.2. Consider a Kerr black hole and two ideally conducting plates at a distance $`L`$ from the hole. The plates are perpendicular to the black hole angular momentum. There is a uniform external magnetic field $`B_0`$ frozen into the plates. The electromagnetic field is force-free. We calculate the force-free magnetosphere in §3.1, formulate the stability problem in §3.2, and show that the magnetosphere can be screw-unstable in §3.3. The results given here constitute a rigorous proof that the magnetosphere of a Kerr black hole is screw-unstable for certain values of $`Lr_+a`$, where $`a`$ is the angular momentum of the hole per unit mass. In fact, this means that a fast, $`ar_+`$, Kerr hole will flare if $`L>Cr_+`$ where $`C`$ is a dimensionless number of order unity. ### 3.1 The force-free magnetosphere We derive the equation for the force-free magnetosphere following the steps similar to Blandford & Znajek (1977). In our case the equations and their derivation are much simpler because the electrostatic potential is everywhere zero due the ideally-conducting plates. The Boyer-Lindquist coordinates are used, with $`c=1`$, and $`G=1`$. The metric is $$ds^2=(1\frac{2Mr}{\mathrm{\Sigma }})dt^2+\frac{4Mar}{\mathrm{\Sigma }}\mathrm{sin}^2\theta dtd\varphi \frac{\mathrm{\Sigma }}{\mathrm{\Delta }}dr^2\mathrm{\Sigma }d\theta ^2\frac{(r^2+a^2)^2\mathrm{\Delta }a^2\mathrm{sin}^2\theta }{\mathrm{\Sigma }}\mathrm{sin}^2\theta d\varphi ^2,$$ (1) $$\mathrm{\Sigma }=r^2+a^2\mathrm{cos}^2\theta ,$$ (2) $$\mathrm{\Delta }=r^22Mr+a^2,$$ (3) $`M`$ is the mass, $`a`$ is the angular momentum per unit mass. The event horizon radius is $`r_+=M+\sqrt{M^2a^2}`$. The magnetosphere is described by the Maxwell equations $$_k(\sqrt{g}F^{ik})=4\pi \sqrt{g}j^i,$$ (4) and the force-free condition $$F_{ik}j^k=0.$$ (5) Here $`g=\mathrm{det}(g_{ik})=\mathrm{\Sigma }^2\mathrm{sin}^2\theta `$, $`F_{ik}=_iA_k_kA_i`$ is the electromagnetic field tensor, $`A^i`$ is the four-potential, $`j^i`$ is the four-current. We consider the stationary axisymmetric case, $`_t=_\varphi =0`$. From (5), $$(j^r_r+ȷ^\theta _\theta )A_t=0,$$ (6) $$(j^r_r+ȷ^\theta _\theta )A_\varphi =0.$$ (7) Then $`A_t`$ must be a function of $`A_\varphi `$. At the plates $`A_t=0`$. Assuming that all the points in the magnetosphere are connected to the conducting plates by the magnetic field lines, $`A_t=0`$ everywhere. Denote $`AA_\varphi `$, $`B_rA_\theta +_\theta A_r`$. Then the remaining two force-free conditions and three of the Maxwell equations <sup>1</sup><sup>1</sup>1The remaining Maxwell equation gives the charge density. We do not need it. are $$j^\varphi _rA+Bj^\theta =0,$$ (8) $$j^\varphi _\theta ABj^r=0,$$ (9) $$_\theta (\sqrt{g}g^{rr}g^{\theta \theta }B)=4\pi \sqrt{g}j^r,$$ (10) $$_r(\sqrt{g}g^{rr}g^{\theta \theta }B)=4\pi \sqrt{g}j^\theta ,$$ (11) $$_r(\sqrt{g}g^{\varphi \varphi }g^{rr}_rA)+_\theta (\sqrt{g}g^{\varphi \varphi }g^{\theta \theta }_\theta A)=4\pi \sqrt{g}j^\varphi .$$ (12) The combination $`\sqrt{g}g^{rr}g^{\theta \theta }B=2I`$, where $`I`$ is the poloidal current. From (10) and (11), $$(j^r_r+ȷ^\theta _\theta )I=0.$$ (13) From (7) and (13), $`I=I(A)`$ \- the current flows along the field lines. From (10), $`2\pi \sqrt{g}j^r=I^{}_\theta A`$, where $`I^{}dI/dA`$. Now from (9), $$j^\varphi =\frac{II^{}}{\pi gg^{rr}g^{\theta \theta }}.$$ (14) With $`j^\varphi `$ given by (14), $`j^r`$ by (10), and $`j^\theta `$ by (11), equations (7)-(11) are satisfied. Equation (12) takes the final form $$_r(g_{00}_rA)+\mathrm{\Delta }^1\mathrm{sin}\theta _\theta ((g_{00}/\mathrm{sin}\theta )_\theta A)+4\mathrm{\Delta }^1\mathrm{\Sigma }II^{}=0,$$ (15) where $`g_{00}=12Mr\mathrm{\Sigma }^1`$. The boundary conditions at the plates are $$A|_{z=\pm L}=B_0\rho ^2/2,$$ (16) where $`zr\mathrm{cos}\theta `$, $`\rho r\mathrm{sin}\theta `$. The boundary condition at the event horizon follows from the requirement that $`F^2`$ be finite at $`r_+`$. This gives $$2I=\frac{a\mathrm{sin}\theta }{\mathrm{\Sigma }}_\theta A|_{r=r_+}.$$ (17) We now solve (15) assuming $`ar_+`$. For $`a=0`$, (17) gives $`I=0`$, and then (15) and (16) give $`A=B_0r^2\mathrm{sin}^2\theta /2`$. We use this $`A`$ in (17) and find the poloidal current, $$I(A)=\frac{aA}{r_+^2}\sqrt{1\frac{2A}{B_0r_+^2}}.$$ (18) We can now use this $`I(A)`$ in (15) and iterate if greater precision is required. But we do not need a more accurate solution for our current purpose. Equation (18) is sufficiently accurate to analyze the stability of the magnetosphere in the limit of $`r_+a`$. ### 3.2 Stability in Force-Free Electrodynamics We have found the stationary magnetic configuration. Now we study its stability in the framework of force-free electrodynamics (FFE) defined by (4), (5). For $`Lr_+`$, we can use the galilean metric. FFE is classical electrodynamics supplemented by the force-free condition. Thus $$_t𝐁=\times 𝐄,$$ (19) $$_t𝐄=\times 𝐁𝐣,$$ (20) $$\rho 𝐄+𝐣\times 𝐁=0.$$ (21) $`𝐁=0`$ is the initial condition. The speed of light is $`c=1`$; $`\rho =𝐄`$ and $`𝐣`$ are the charge and current densities multiplied by $`4\pi `$. The electric field is everywhere perpendicular to the magnetic field, $`𝐄𝐁=0`$ <sup>2</sup><sup>2</sup>2If $`\rho 0`$, $`𝐄𝐁=0`$ follows from (21), if $`\rho =0`$, this condition is an independent basic equation of FFE.. The electric field component parallel to the magnetic field should vanish because the charges are freely available in FFE. It is also assumed that the electric field is everywhere weaker than the magnetic field, $`E^2<B^2`$. Then equation (21) means that it is always possible to find a local reference frame where the field is a pure magnetic field, and the current is flowing along this field. FFE is Lorentz invariant. Equation (21) can be written in the form of the non-linear Ohm’s law $$𝐣=\frac{(𝐁\times 𝐁𝐄\times 𝐄)𝐁+(𝐄)𝐄\times 𝐁}{B^2}.$$ (22) Equations (19), (20), (22) form an evolutionary system (initial conditions $`𝐁=0`$ and $`𝐄𝐁=0`$ are implied). It therefore makes sense to study stability of equilibrium electromagnetic fields in FFE. One can also study linear waves and their nonlinear interactions in the framework of FFE (Thompson & Blaes 1998). It is convenient to introduce a formulation of FFE similar to magnetohydrodynamics (MHD); then we can use the familiar techniques of MHD to test stability of magnetic configurations. To this end, define a field $`𝐯=𝐄\times 𝐁/B^2`$, which is similar to velocity in MHD. Then $`𝐄=𝐯\times 𝐁`$ and equation (19) becomes the “frozen-in” law $$_t𝐁=\times (𝐯\times 𝐁).$$ (23) From $`𝐯=𝐄\times 𝐁/B^2`$, and from equations (19)-(21), one obtains the momentum equation $$_t(B^2𝐯)=\times 𝐁\times 𝐁+\times 𝐄\times 𝐄+(𝐄)𝐄,$$ (24) where $`𝐄=𝐯\times 𝐁`$. Equations (23), (24) are the usual MHD equations except that the density is equal $`B^2`$ and there are order $`v^2`$ corrections in the momentum equation. A force-free magnetic field ($`\times 𝐁\times 𝐁=0`$) with a zero electric field is a stationary FFE solution. We will study the stability of the force-free magnetic configurations using the MHD formulation of FFE. Linear perturbations, denoted $`𝐛`$ and $`𝐯`$, $`\mathrm{exp}(i\omega t)`$, satisfy $$i\omega 𝐛=\times (𝐯\times 𝐁),$$ (25) $$i\omega B^2𝐯=\times 𝐁\times 𝐛+\times 𝐛\times 𝐁.$$ (26) Define the displacement $`𝝃`$ by $`𝐯=i\omega 𝝃`$. From (25), $`𝐛=\times (𝝃\times 𝐁)`$. Now equation (26) can be written as $$\omega ^2B^2𝝃=\times 𝐁\times \times (𝝃\times 𝐁)+\times \times (𝝃\times 𝐁)\times 𝐁\widehat{𝐊}𝝃.$$ (27) Since the operator $`\widehat{𝐊}`$ is self-adjoint (Kadomtsev 1966 and references therein), the frequency is given by the variational principle $$\omega ^2=\mathrm{min}\frac{d^3r𝝃\widehat{𝐊}𝝃}{d^3rB^2𝝃^2}.$$ (28) The equilibrium field $`𝐁`$ is unstable if the potential energy $$Wd^3r𝝃\widehat{𝐊}𝝃,$$ (29) is negative for some displacements $`𝝃`$. Consider the case of cylindrical symmetry, coordinates $`(r,\theta ,z)`$. The equilibrium field $`𝐁=(0,U(r),B(r))`$ should satisfy $`BB^{}+Ur^1(rU)^{}=0`$, where the prime denotes the r-derivative. For an eigenmode $`\mathrm{exp}(im\theta +ikz)`$, the potential energy reduces to (Kadomtsev 1966) $$W𝑑r(f\xi ^2+g\xi ^2),$$ (30) where $`\xi `$ is the radial component of the displacement. The other two components of the displacement vector, $`\xi _\theta `$ and $`\xi _z`$, were chosen to minimize the energy for a given $`\xi `$. The functions of radius $`f`$ and $`g`$ are given by $$f=r\frac{(krB+mU)^2}{k^2r^2+m^2},$$ (31) $$g=r^1\left(\frac{k^2r^2+m^21}{k^2r^2+m^2}(krB+mU)^2+\frac{2k^2r^2}{(k^2r^2+m^2)^2}(k^2r^2B^2m^2U^2)\right).$$ (32) We now show that magnetic configurations with a non-zero toroidal field $`U`$ are screw unstable. Screw means, e. g., that $`m=1`$, but not $`m=1`$, is unstable for a given sign of $`k`$. Kadomtsev (1966) gives a clear discussion of the screw instability in plasmas, and shows that the screw mode is the most dangerous mode (if the plasma is unstable, it is screw unstable). Assume that $`U(0)=0`$, $`U^{}(0)>0`$, $`U0`$ for $`r\mathrm{}`$, and $`U`$ is positive in between. Assume that $`B`$ is everywhere positive. Let $`k`$ be positive and small. Take $`m=1`$. Let $`r_0`$ be the first zero of $`krBU`$. Take a trial function $`\xi =1`$ for $`r<r_0`$, and zero at $`r>r_0`$. We can chose this generalized function $`\xi `$ in such a way that the first term in the energy integral (31) vanishes. The second term is an integral of $`g`$ from $`0`$ to $`r_0`$. For small $`k`$, $`g=k^2r(krBU)(3krB+U)`$ is negative. Thus an eigenmode $`\mathrm{exp}(i\theta +ikz)`$ is screw-unstable if $`krBU`$. If the length of the magnetosphere in the z-direction is $`L`$, the minimal value of $`k`$ is $`L^1`$. The instability condition (the Kruskal-Shafranov criterion) is $$\frac{U}{r}>C\frac{B}{L},$$ (33) where $`C`$ is a dimensionless number of order unity. The value of $`C`$ depends on the boundary conditions. ### 3.3 Stability of the Black Hole Magnetosphere For the black hole magnetosphere, in the limit $`r_+a`$, at a distance from the hole $`r_+`$, the poloidal field $`B`$ is approximately constant and equal to $`B_0`$. The poloidal current is given by (18), and the toroidal field is $$U=\frac{a}{r_+}B_0\frac{r}{r_+}\sqrt{1\frac{r^2}{r_+^2}},$$ (34) in the cylindrical coordinates of §3.2. According to the Kruskal-Shafranov criterion, the magnetosphere will be screw unstable if $`L>Cr_+^2/a`$, where $`C`$ is a dimensionless number of order unity. ## 4 Discussion The flaring is a new mechanism for the extraction of the rotational energy of Kerr holes. Many previous studies considered the Blandford-Znajek mechanism in which the magnetic field connects the event horizon to infinity. By contracts, the flaring occurs when the magnetic field ends in the accretion disk. Both mechanisms give the same averaged luminosity, as expected from dimensional analysis. The essential feature of our mechanism is that the rotational frequency of the hole determines the time structure of the radiated energy. The hole frequency may appear observationally as a quasi-periodicity of the X-ray flux. It would be interesting to search for this periodicity in the high-time-resolution data. ###### Acknowledgements. I thank John Bahcall for discussions. This work was supported by NSF PHY-9513835.
no-problem/9908/hep-ph9908474.html
ar5iv
text
# Lattice matrix elements confronting the experimental value of 𝜀'/𝜀 ## Abstract: A new lattice estimate of $`K2\pi `$ transitions claims, contrary to all other computations, that the hadronic matrix element for the gluon penguin operator $`Q_6`$ has opposite sign and, in addition, is much larger than the vacuum saturation approximation. We comment under what conditions (if any) it is possible to reconcile this lattice result with the experimental value of $`\epsilon ^{}/\epsilon `$ . The dramatic impact of new physics in the kaon system that seems to be required is not easily accommodated within our present theoretical understanding. Kaon Physics, CP violation, Lattice QCD preprint: SISSA 102/99/EP August 1999 1. A new lattice estimate of $`K2\pi `$ transitions using domain-wall fermions claims , contrary to all other computations, that the hadronic matrix element for the gluon penguin operator $`Q_6`$ has opposite sign and, in addition, is much larger than its vacuum saturation approximation. This surprising result comes about because of the contribution of the so-called eye-contraction diagrams. The resulting value for the CP violating parameter $`\epsilon ^{}/\epsilon `$ is of the opposite sign and almost one order of magnitude bigger than the current experimental determination: $$\epsilon ^{}/\epsilon =(2.1\pm 0.46)\times 10^3,$$ (1) which is obtained by averaging over the KTeV and NA48 preliminary results as well as the older 1992-93 experiments (E731 and NA31 ). If this lattice result (taken at its face value) will stand further scrutiny it raises serious questions on our understanding of electroweak physics in the kaon system. In this letter we would like to discuss under what conditions (if any) it is possible to reconcile a large and positive $`2\pi |Q_6|K`$ with the experimental value of $`\epsilon ^{}/\epsilon `$ . We consider two possible scenarios: * A modification of short-distance physics that changes both the sign and the size of the Wilson coefficient of the gluon or electroweak penguin operators. This can in principle be achieved by tampering with the initial matchings of the various coefficients while preserving the standard basis of operators. As we shall discuss, the behavior of the renormalization group equations (RGE) for the relevant effective lagrangian force us to rather extreme changes in order to achieve the desired effect. Even though supersymmetry provides a framework for potentially large effects, to construct a model in which such large deviations from standard physics are present in $`\epsilon ^{}`$ while conspiring to be invisible everywhere else involves a contrived set of assumptions. * An enlargement of the standard operator basis. We limit our discussion to the case of the chromomagnetic penguin operator because it does not affect the renormalization of the Wilson coefficients of the other dimension six operators (allowing us to draw model-independent conclusions) and it has been already shown to be a promising candidate for new-physics effects in $`\epsilon ^{}/\epsilon `$ . Given the current estimate of this operator’s matrix element, short-distance changes alone—although potentially very large—are not sufficient in bringing $`\epsilon ^{}/\epsilon `$ in the experimental ball-park. A final assessment requires a lattice evaluation of the hadronic matrix element of the chromomagnetic operator. Both scenarios call for a dramatic impact of new physics in the kaon system and they are not easily accommodated within our present theoretical understanding. More exotic extensions of the standard operator basis and more extreme scenarios require a detailed model-dependent discussion which is beyond the scope of the present letter. <sup>1</sup><sup>1</sup>1Possible effects of dissipation and loss of coherence in the kaon system relevant to $`\epsilon ^{}/\epsilon `$ have been discussed in ref. . 2. Let us fix our notation by introducing the $`\mathrm{\Delta }S=1`$ effective four-quark lagrangian for $`m_c<\mu <m_b`$ $$_{eff}^{\mathrm{\Delta }S=1}=\frac{G_F}{\sqrt{2}}\lambda _u\left\{(1\tau )\underset{i=1,2}{}C_i(\mu )\left[Q_i(\mu )Q_i^c(\mu )\right]+\tau \underset{i=1,10}{}C_i(\mu )Q_i(\mu )\right\},$$ (2) where $`\lambda _qV_{qd}V_{qs}^{}`$ and $`\tau =\lambda _t/\lambda _u`$. In the discussion that follows we will assume as the standard model (SM) reference value for the CP phase $`\text{Im}\lambda _t=10^4`$. The renormalization group Wilson coefficient $`C_i(\mu )`$ are known to the next-to-leading order in $`\alpha _s`$ and $`\alpha _e`$ . The standard basis of effective operators is $$\begin{array}{cccc}\hfill Q_1& =& \left(\overline{s}_\alpha u_\beta \right)_{\mathrm{V}\mathrm{A}}\left(\overline{u}_\beta d_\alpha \right)_{\mathrm{V}\mathrm{A}},\hfill & Q_1^c=\left(\overline{s}_\alpha c_\beta \right)_{\mathrm{V}\mathrm{A}}\left(\overline{c}_\beta d_\alpha \right)_{\mathrm{V}\mathrm{A}},\hfill \\ \hfill Q_2& =& \left(\overline{s}u\right)_{\mathrm{V}\mathrm{A}}\left(\overline{u}d\right)_{\mathrm{V}\mathrm{A}},\hfill & Q_2^c=\left(\overline{s}c\right)_{\mathrm{V}\mathrm{A}}\left(\overline{c}d\right)_{\mathrm{V}\mathrm{A}},\hfill \\ \hfill Q_{3,5}& =& \left(\overline{s}d\right)_{\mathrm{V}\mathrm{A}}_q\left(\overline{q}q\right)_{\mathrm{V}\mathrm{A}},\hfill & \\ \hfill Q_{4,6}& =& \left(\overline{s}_\alpha d_\beta \right)_{\mathrm{V}\mathrm{A}}_q(\overline{q}_\beta q_\alpha )_{\mathrm{V}\mathrm{A}},\hfill & \\ \hfill Q_{7,9}& =& \frac{3}{2}\left(\overline{s}d\right)_{\mathrm{V}\mathrm{A}}_q\widehat{e}_q\left(\overline{q}q\right)_{\mathrm{V}\pm \mathrm{A}},\hfill & \\ \hfill Q_{8,10}& =& \frac{3}{2}\left(\overline{s}_\alpha d_\beta \right)_{\mathrm{V}\mathrm{A}}_q\widehat{e}_q(\overline{q}_\beta q_\alpha )_{\mathrm{V}\pm \mathrm{A}},\hfill & \end{array}$$ (3) where $`V\pm A`$ stands for $`\gamma _\mu (1\pm \gamma _5)`$ and $`\widehat{e}_q`$ is the value of the electric charge of the quark $`q=u,d,s,c`$. 3. Taking the hadronic matrix elements of the operators in eq. (3) obtained by the lattice calculation , and given their standard model Wilson coefficients, $`\epsilon ^{}/\epsilon `$ is completely dominated by the contribution of $`Q_6`$ and is therefore large and negative (in dramatic disagreement with the experiment). In order to reconcile the lattice and the experimental result, we must somehow compensate for this dominant contribution. The ratio $`\epsilon ^{}/\epsilon `$ is determined by the sum of isospin $`I=0`$ and $`I=2`$ amplitudes. Let us examine in turns possible modifications on these two classes of contributions. A first possibility, in the $`I=0`$ amplitude, is that the Wilson coefficient of the $`Q_6`$, at the scale at which the matrix element is computed (about 2 GeV), is changed with respect to its standard model value in order to reproduce the experimental result. Assuming that new physics only modifies the initial conditions (at $`m_W`$) of the RGE, we need to know how these changes are propagated by the RGE down to the hadronic scale. This has been discussed for the whole operator basis in section VII of ref.. In order to keep the discussion as model independent as possible, we parameterize all deviations from the SM matching conditions $`C_i(m_W)`$ in terms of the parameters $$r_iC_i(m_W)/C_i^{SM}(m_W).$$ (4) In general, a Wilson coefficient at the low scale receive contributions from both a multiplicative and additive renormalization, the latter arising from QCD-induced operator mixing. In particular, the $`C_6(\mu )`$ is dominated in the standard basis by the additive renormalization induced by the mixing of $`Q_6`$ with the $`Q_2`$ operator. We have plotted in Fig. 1 the Wilson coefficient $`C_6`$(2 GeV) as a function of the parameters $`r_2`$ and $`r_6`$. The dark (red) band represents the values of $`C_6`$ for which the experimental result for $`\epsilon ^{}/\epsilon `$ would be recovered (leaving all other Wilson coefficients unchanged), that is $$0.2<\frac{C_6(2\text{GeV})}{C_6^{SM}(2\text{GeV})}<0.1.$$ (5) As it can seen by inspection, this possibility is realized either by a drastic reduction of $`C_2(m_W)`$ (the left side of Fig. 1) or a large enhancement of $`C_6(m_W)`$ (the right side of Fig. 1). However, the value of $`C_2`$(2 GeV) is severely constrained by the CP-conserving amplitude $`A_0(K2\pi )`$, which in the same lattice calculation is reproduced up to a factor of two. Moreover, it is difficult to substantially change $`C_2(m_W)`$ by means of new physics because this Wilson coefficient is due to tree level $`W`$-exchange. Looking then at Fig. 1, and given that $`C_2(m_W)`$ cannot differ too much from its standard-model value, it is possible to reproduce the experimental $`\epsilon ^{}/\epsilon `$ only by enhancing $`C_6(m_W)`$ by more than a factor of twenty. However, such a large enhancement can hardly take place without affecting other processes and we shall come back to it after discussing the electroweak penguin. Leaving the $`I=0`$ contribution alone, we can still compensate for the large and negative result by acting on the the $`I=2`$ contribution. Here the dominant operator is $`Q_8`$. Given the lattice estimate of the matrix element of this operator, agreement with the experiments would require $$C_8(2\text{GeV})30C_8^{SM}(2\text{GeV}).$$ (6) The RGE analysis shows that $`C_8`$(2 GeV) varies proportionally to $`C_7(m_W)`$ (recall that $`C_8(m_W)`$=0) and, as one can see from Table 1, $`C_7(m_W)`$ receives contributions from photon and $`Z`$ penguins. To date, the best limits on the CP conserving component of the $`Z`$-penguin operator, $`\text{Re}𝒫_Z`$, are provided by the $`K_L\mu ^+\mu ^{}`$ decay , whose branching ratio is measured to be $$B(K_L\mu ^+\mu ^{})=(7.18\pm 0.17)\times 10^9,$$ (7) even though the constraint on $`\text{Re}𝒫_Z`$ is not as accurate as the experimental precision because of the theoretical long-distance uncertainties related to the two-photon component . The cleanest constraint on $`|𝒫_Z|`$ comes from the decay $`K^+\pi ^+\overline{\nu }\nu `$, which is currently measured to be $$B(K^+\pi ^+\overline{\nu }\nu )=4.2_{3.5}^{+9.7}\times 10^{10}.$$ (8) Due to the lack of further evidence, the branching ratio in eq. (8) is going to be reduced by a factor of two or three . Taking the standard model expectation $`B(K^+\pi ^+\overline{\nu }\nu )=(0.8\pm 0.3)\times 10^{10}`$ as a reference point, a numerical analysis shows that we can at most modify $`\text{Im}𝒫_Z^{SM}`$ by a factor of sixteen. Barring the unlikely possibility of independent and widely different effects in the effective $`(\overline{s}d)_{VA}`$ vertex of the photon and $`Z`$ penguins, eq. (8) rules out the enhancement required by eq. (6). Analogous considerations hold for the other electroweak operators which would entail even larger deviations of their Wilson coefficients from the standard model values. By the same token, also the enhancement by a factor of twenty of the gluonic penguin coefficient $`C_6(m_W)`$—which we argued is necessary in the $`I=0`$ amplitude in order to reproduce the experimental value of $`\epsilon ^{}/\epsilon `$ —is difficult to accommodate. A notable exception to the above argument is the presence of gluino-induced flavour-changing neutral currents in supersymmetric models. The gluino induced $`\mathrm{\Delta }S=1`$ transitions are suppressed in Z-penguin diagrams compared to gluon or photon penguins by a factor of $`O(m_K^2/m_Z^2)`$ (for a detailed discussion of this effect which follows from gauge invariance see ref. ). As a consequence, the bounds on Z-penguins are not effective on gluino-induced gluon and photon penguins. However, the gluino contributions to the standard $`\mathrm{\Delta }S=1`$ penguin operators are indirectly constrained by the sharp bounds on the gluino-induced $`\mathrm{\Delta }S=2`$ box diagrams relevant to $`\overline{K}^0K^0`$ transitions . Gluino-box diagrams can also play a direct role in $`\mathrm{\Delta }S=1`$ amplitudes. In ref. , it is shown that gluino-box diagrams may induce a potentially large isospin-breaking contribution to the electroweak penguin $`Q_8`$ while satisfying all other bounds. However, the large factor required by eq. (6) implies a rather large mass splitting between the right-handed squark isospin doublet ($`m_{\stackrel{~}{u}_R}m_{\stackrel{~}{d}_R}1`$ TeV) together with a large gluino mass in order to evade the bounds from $`\mathrm{\Delta }m_K`$ and $`\epsilon `$. 4. An operator not usually included in the standard model analysis of $`\epsilon ^{}/\epsilon `$ is the chromomagnetic penguin $$Q_{11}=\frac{g_s}{8\pi ^2}\overline{s}\left[m_dR+m_sL\right]\sigma _{\mu \nu }G^{\mu \nu }d,$$ (9) where $`R(L)=(1\pm \gamma _5)/2`$ and $`G^{\mu \nu }`$ is the gluon field. The matrix element $`2\pi |Q_{11}|K`$ has been computed in the chiral quark model and shown to arise only at $`O(p^4)`$ in the chiral expansion and to be further suppressed by a $`m_\pi ^2/m_K^2`$ factor with respect to the naive expectation . For these reasons, even though the $`Q_{11}`$ Wilson coefficient receives a large additive renormalization, its standard model contribution to $`\epsilon ^{}/\epsilon `$ has been shown to be negligible . This is no longer true if the Cabibbo-Kobayashi-Maskawa (CKM) suppression of its CP violating component can be lifted without violating other bounds. A clever example of it is discussed in ref. , where the standard model factor $`\text{Im}\lambda _tO(\mathrm{sin}^5\theta _C)`$ is replaced in a supersymmetric framework by a CP violating squark mixing of $`O(\mathrm{sin}\theta _C)`$, thus introducing a potential enhancement of the chromomagnetic Wilson coefficient by three orders of magnitude. This solution allows for a contribution to $`\epsilon ^{}/\epsilon `$ at the $`10^3`$ level by saturating the known bounds coming from CP violating phenomenology (discussions of the implications of these bounds on various supersymmetric models can be found in ref. ). This enhancement of $`C_{11}`$ is still not enough to compensate for the huge negative $`Q_6`$ contribution to $`\epsilon ^{}/\epsilon `$ of the lattice result; in fact, keeping fixed the gluon penguin contribution, agreement with experiment would require a matrix element for $`Q_{11}`$ larger by about a factor of ten. Actually, the leading $`O(p^4)`$ chiral quark model estimate of the hadronic matrix element may receive potentially large $`O(p^6)`$ contributions if the accidental $`m_\pi ^2/m_K^2`$ suppression is replaced by $`m_K^2/\mathrm{\Lambda }_\chi ^2`$. To further assess this possibility, it would therefore be interesting to have an estimate of the $`Q_{11}`$ matrix element from the same lattice approach that produces the large and positive matrix element for the $`Q_6`$. Clearly, extensions of the standard model can also involve new effective operators beyond the standard basis of eq. (3) and $`Q_{11}`$. However, in this case very little can be said without a complete re-analysis of $`\epsilon ^{}/\epsilon `$ . Similarly, scenarios in which the CKM matrix is taken to be real and CP violation arises only in the new-physics sector can in principle be invoked but again require a detailed model-dependent analysis before being considered a viable alternative. 5. In conclusion, while a combination of the above scenarios may make some of the requirements less severe, a viable model which avoids the phenomenological bounds discussed is forced to rely on a contrived choice of parameters. It is nevertheless remarkable how supersymmetry can provide a framework for potentially large effects on $`\epsilon ^{}/\epsilon `$ while satisfying all present data. It is also fair to add that the lattice result by means of domain-wall fermions in ref. must stand further scrutiny and corroboration before concluding that the standard model is facing its most dramatic challenge to date. We thank F. Borzumati and M. Neubert for useful discussions and comments.
no-problem/9908/astro-ph9908287.html
ar5iv
text
# The Very Extended Ionized Nebula around the Quasar MR2251-178 ## 1 Introduction In the ongoing quest to better understand the luminous quasar population, one of the most fertile areas of study has been the investigation of their galaxy environments. The host galaxy not only must provide fueling for the central engine, but should also display the effects of the strong nuclear ionizing emission and the impact of dynamic activity, such as jets and winds. Broadband observations of quasar environments suggest a surprising variety of host galaxy morphologies, including spirals and ellipticals, nearby companions, and tidal interactions. While early ground-based work indicated a large percentage of spirals amongst low-redshift quasar hosts ($`40`$%; Hutchings et al. 1984a ) and no concrete evidence for ellipticals, subsequent HST observations find an elliptical fraction of greater than half (Bahcall et al. (1997)). More recent HST work suggests in fact that nearly all radio-loud and radio-quiet quasars reside in massive ellipticals (McLure et al. (1999); see also Disney et al. (1995)). A substantial fraction of quasar host galaxies exhibit tidal tails and streamers indicative of interactions (e.g., Hutchings & Campbell (1983); Hutchings et al. 1984b ; Stockton & MacKenty (1987); Bahcall et al. (1997); Hutchings et al. (1999)). Evidence for extended gas from interactions has also been found at H I 21 cm (e.g., Lim & Ho (1999)). The presence of companion galaxies is common; the QSO/galaxy correlation function is evidently several times that for galaxies alone (Fisher et al. (1996)). Given the strong line emission in the nuclei of active galaxies and quasars, as well as the prevalence of ionized gas in interacting systems, narrowband observations of quasar environments have proved interesting as well. The imaging survey of Stockton & MacKenty (1987), the largest to date, found highly-structured \[O III\] emission in a quarter of 47 luminous QSOs, with typical extents of a few tens of kpc. Similar spectroscopic observations have detected extended regions of line emission in half of the objects observed (Boroson et al. (1985)). The line ratios usually suggest photoionization by the nuclear power-law spectrum (e.g., Boroson et al. (1985); Boisson et al. (1994)), although stellar absorption lines from (presumably in-situ) stars are sometimes seen (e.g., Miller et al. (1996)). Kinematic studies of the ionized gas component are scarce, but the gas motions generally appear to be complex and chaotic (e.g., Durret et al. (1994)). The size and luminosity of the ionized component appear to be correlated with both the narrow-line nuclear luminosity and the radio power of the quasar (Boroson et al. (1985); Stockton & MacKenty (1987); Véron-Cetty & Woltjer (1990)). The quasar MR 2251–178 is one of a few radio-quiet quasars which exhibit an extended gaseous envelope (e.g., Bergeron et al. (1983)). This quasar was first discovered as a strong and variable X-ray source by the Ariel V satellite (Cooke et al. (1978)). Subsequent observations identified the object as a quasar at a redshift of $`0.0638\pm 0.0015`$ (Ricker et al. (1978); Canizares et al. (1978)), residing in the outskirts of a small cluster (Phillips (1980)). In this Letter, we present deep H$`\alpha `$ observations of MR 2251–178 obtained with the TAURUS Tunable Filter (TTF), a new, etalon-based instrument which has been optimized for the detection of faint, extended emission-line gas. These new data allow us to better constrain the extent, velocity field, and origin of the ionized nebula around MR 2251–178. ## 2 Observations and Reductions MR 2251–178 was observed on August 30, 1998 using the TTF at the f/8 Cassegrain focus of the 3.9-meter AAT. The TTF instrument consists of a pair of modified high-finesse ($`N40`$) Queensgate etalons (blue and red) which can be tuned to provide narrowband imaging anywhere within the wavelength range 400 to 960 nm, through an arbitrary bandpass, with resolving powers of 100 to 1000.<sup>1</sup><sup>1</sup>1Consult Bland-Hawthorn & Jones (1998) and the TTF web page (http://www.aao.gov.au/local/www/jbh/ttf/) for further details. The observations of MR 2251–178 were made with the red side of the TTF, using a mediumband ($`\mathrm{\Delta }\lambda =26.0`$ nm) blocking filter centered at 707 nm, tilted by 16°. Two 600-second exposures were obtained at each of two etalon spacings. The exposures were dithered amongst pointings on a 15″ grid. The average atmospheric seeing of 1.3″ was oversampled by the 0.37″ pixels. The night was photometric. MR 2251–178 was also observed on September 3, 1998 in direct imaging mode, using a standard I-band filter. Fits to a number of emission lines from observations of a calibration lamp (CuAr) were used to determine the relationship between wavelength, spatial position, and etalon gap spacing. The free spectral range of the etalon, i.e., the distance between orders, was found to be 265.8 Å, well-matched to the bandwidth of the blocking filter. The system was used in the 26th order of interference, with a 9 micron etalon gap, at an effective finesse of 39.7. This translates to a spectral resolution of 6.8 Å, or $`R1040`$, with an effective bandpass of $`12`$ Å. The pair of etalon spacings produced imagery with central wavelengths on the optical axis of 6983 Å and 6986 Å, corresponding to redshifted velocities of 60 and 195 km s<sup>-1</sup> relative to the quasar. The field is essentially monochromatic: pixels $`1`$′ from the optical axis have a central wavelength only 1.3 Å ($`60`$ km s<sup>-1</sup>) bluer than pixels on-axis. The optical axis is located approximately 35″ northwest of the quasar. The data frames were bias-subtracted and flatfielded in the typical manner. An azimuthally-symmetric sky frame was produced from a mean radial sky spectrum for each image and subsequently subtracted. Images at each of the two etalon spacings were then aligned and combined using 9 stellar objects in the field. An image mask was used to simultaneously remove three ghost reflections from each image. The data were flux calibrated using observations of the standard star EG 21 (Stone & Baldwin (1983)). The I-band observations were reduced following standard CCD procedures. The summed continuum image was matched to the narrowband imagery using a number of stars and subsequently subtracted. A few remaining stellar residuals and cosmetic defects were identified and repaired by hand. ## 3 Discussion Figure The Very Extended Ionized Nebula around the Quasar MR2251-178 presents our final H$`\alpha `$ images of MR 2251–178 at each of the two etalon spacings (panels $`a`$ and $`b`$), as well as the I-band continuum image (panel $`c`$) and a combined H$`\alpha `$ image (panel $`d`$). The 2-$`\sigma `$ detection limit of the H$`\alpha `$ images is $``$ 5 $`\times `$ 10<sup>-18</sup> erg s<sup>-1</sup> cm<sup>-2</sup> arcsec<sup>-2</sup>, more than an order of magnitude fainter than previously published data on this object. In the discussion that follows, we assume $`H_0=50`$ km s<sup>-1</sup> Mpc<sup>-1</sup> and a corresponding image scale of 1.9 kpc arcsec<sup>-1</sup>. ### 3.1 Morphology Published optical imagery and spectroscopy have detected two ionized gas components around the quasar MR 2251–178: an elongated, highly-ionized circumnuclear component of diameter $`27`$ kpc, and an extended envelope of faint H$`\alpha `$\- and \[O III\]-emitting filaments out to a radius of $`110`$ kpc (Bergeron et al. (1983); Hansen et al. (1984); Alighieri et al. (1984); Macchetto et al. (1990)). Our observations confirm these findings and reveal a number of new features. The circumnuclear component of ionized gas is evidenced by strong H$`\alpha `$ emission directly surrounding the quasar, extending out to a radius of $`20`$ kpc. This emission is slightly elongated in the east-west direction, i.e., along the axis of the quasar’s radio jet ($`PA102`$°; Macchetto et al. (1990)), and has been identified as an “extended emission-line region” (EELR) of the quasar host galaxy (Mulder & Valentijn (1992)). As others have reported on characteristics of the EELR in some detail, we will say nothing more of it herein. The extended ionized gas component is observed as a much larger network of diffuse and filamentary emission exterior to the EELR. Above a flux level of $`1.8\times 10^{17}`$ erg s<sup>-1</sup> cm<sup>-2</sup> arcsec<sup>-2</sup>, diffuse H$`\alpha `$ emission is visible surrounding the galaxy to a radius of at least 50 kpc. Knots and filaments of ionized gas are visible to radii up to 120 kpc towards the northeast, southeast, and southwest. The filaments display a spiral morphology and exhibit a remarkable azimuthal symmetry that has not been previously seen. The total extent of this complex ($`200`$ kpc) makes it the largest known around a quasar. We derive a total flux from the extended ionized gas component of $`1.4\times 10^{13}`$ erg s<sup>-1</sup> cm<sup>-2</sup>, or a luminosity of $`2.5\times 10^{42}`$ erg s<sup>-1</sup> at the redshift distance of the quasar (384 Mpc). This value, as well as the corresponding upper limit on the gas mass of $`6\times 10^{10}`$ M, is comparable or slightly greater than previously published results (e.g., Nørgaard-Nielsen et al. (1986); Macchetto et al. (1990)). A few of the individual knots and filaments have been previously noted; these are indicated in Figure The Very Extended Ionized Nebula around the Quasar MR2251-178$`d`$, using the denotations from Macchetto et al. (1990). Several additional knots also noted by these authors and others (e.g., Alighieri et al. (1984)) are conspicuous in their absence from our imagery. Most notable among these are the “E” knots interior to the eastern filament in Figure The Very Extended Ionized Nebula around the Quasar MR2251-178. As will be noted in the following section, this absence is probably due to the large range of velocity spanned by the filament emission, relative to our narrow bandpasses. ### 3.2 Kinematics The spectroscopy which initially discovered the extended ionized component of MR 2251–178 also provided the first kinematical information on the nebula (Bergeron et al. (1983)). Those observations revealed a normal rotation curve along each of three position angles, with peak velocities of $`150`$ km s<sup>-1</sup>. Surprisingly, the extended nebula appears to be rotating in an opposite sense to the inner (EELR) regions of the galaxy (e.g., Nørgaard-Nielsen et al. (1986)). The only published two-dimensional kinematics of the system are narrow-field H$`\alpha `$ Fabry-Perot observations of the EELR (Mulder & Valentijn (1992)). Comparison of our H$`\alpha `$ imagery at each of the two distinct etalon positions reveals a broad velocity gradient across the quasar nebula. The maximum gradient appears to be along a position angle of $`40`$°, with the eastern filaments generally less redshifted than the western filaments. The eastern filaments are clearly detected in both the “blue” and “red” images. Since an upper limit of $`200`$ km s<sup>-1</sup> has been measured for their line widths (Bergeron et al. (1983)), the central velocity of these filaments must lie within the overlap region of our two velocity regimes. Accounting for its distance from the optical axis, the ionized gas east of the quasar must be near the systemic velocity of the quasar. The western filaments are detected only in the “red” image and appear to be redshifted from the systemic velocity by a value of $`200300`$ km s<sup>-1</sup>. This correlates well with the imaging observations of Macchetto et al. (1990), whose broad ($`\mathrm{\Delta }\lambda =64`$ Å) H$`\alpha `$ filter barely detected the western filaments out to a velocity of $`200`$ km s<sup>-1</sup> relative to the quasar. Furthermore, these authors more readily detected the western filaments in their \[O III\] imagery, which encompassed a redder range of velocities. We therefore confirm that the velocity structure of the extended ionized gas component does not appear to follow the rotation curve of the quasar (Nørgaard-Nielsen et al. (1986); Mulder & Valentijn (1992)). The broad velocity range of our images and the limitation of two distinct velocity samples restricts us from analyzing more precise velocity variations, e.g., amongst individual knots and filaments. ### 3.3 The Origin of the H II Envelope Several scenarios have been put forth for the origin of the ionized gas envelope in MR 2251–178, including tidal debris from an interaction with galaxy G1 some $`2\times 10^8`$ years ago, similar debris from a merging event, clouds expelled from the quasar and/or host galaxy, a cooling flow, or the ionized portion of a large H I envelope that is gravitationally bound to the quasar (e.g., Nørgaard-Nielsen et al. (1986); Macchetto et al. (1990)). Our observations reveal a tremendous spatial extent for the emission in this system. We also observe a relatively high degree of symmetry in the envelope, including at least two “arms” of ionized gas and a pervasive diffuse component. This symmetry, together with the organized large-scale kinematics, casts doubt upon the interaction model for creation of the ionized gas. Furthermore, the galaxy G1, the usual culprit quoted as an interacting companion to MR 2251–178, has a velocity redshifted by 1246 km s<sup>-1</sup> relative to the quasar (Bergeron et al. (1983)), yet its spatial position corresponds to that of relatively blueshifted filaments. Published spectroscopy shows the emission lines from these filaments to be symmetric and significantly narrower than this (e.g., Bergeron et al. (1983)), although we cannot rule out the presence of a very faint, high-velocity streamer extending out toward G1. Nevertheless, the coarse kinematic structure of the extended ionized gas indicates that it represents a distinct component from the inner EELR region, since the latter appears to rotate in an opposite sense. This implies that the envelope most probably did not originate within the host galaxy of the quasar. A retrograde merging event, in which a small, gas-rich galaxy has been subsumed by the quasar, remains a possibility, however again the relatively well-ordered large-scale kinematics and azimuthal symmetry of the ionized envelope are problematic. The cooling flow hypothesis appears equally unlikely. As pointed out by Macchetto et al. (1990), the large size and ordered kinematics of the envelope, and the off-centered position and peculiar systemic velocity of the quasar with respect to the underlying cluster, all argue against this scenario. By elimination, our observations therefore favor a model in which the extended ionized envelope resides within a large complex of H I gas centered about the quasar. If this is correct, a deep search for 21 cm H I line emission should reveal the massive neutral envelope and constrain its origin (e.g., captured intergalactic H I clouds, remnant accreting gas from galaxy formation, etc.). ### 3.4 The Ionization of the H II Envelope The high excitation level of the brighter knots around MR 2251–178 (e.g., Bergeron et al. (1983); Nørgaard-Nielsen et al. (1986); Macchetto et al. (1990)) requires an energetic source of ionization. In-situ ionization by a faint hot stellar component is unlikely to be dominant for a number of reasons. Deep broadband imagery of MR 2251–178 reveals a morphology very unlike that of our narrowband observations. The R-band image of Hutchings et al. (1999) exhibits a faint component extended along the N–S direction, markedly different from the symmetric spiral pattern that we observe in H$`\alpha `$. Moreover, Hutchings et al. (1999) do not detect an excess of continuum emission at the locations of any of the brighter H$`\alpha `$ knots in the envelope. Perhaps even more damaging to the stellar ionization scenario is our detection of a large amount of diffuse ionized gas outside of the brighter knots. We therefore favor an external source of ionization. The presence of a significant amount of ionized gas at large angles relative to the jet axis (PA $`102`$°; Macchetto et al. (1990)) seems to rule out ionization mechanisms directly associated with the quasar jet (e.g., shocks). We are therefore left with the possibility that the ionization of the nebula is sustained by the quasar radiation field. As described in detail in previous studies (e.g., Bergeron et al. (1983); Nørgaard-Nielsen et al. (1986); Macchetto et al. (1990)), the power radiated by the quasar can easily account for the high ionization level of the nebula. However, the relative symmetry of the envelope requires that the ionizing radiation is escaping the quasar symmetrically with respect to our line of sight. The large-scale radiation field from MR 2251–178 therefore shows no sign of anisotropy or alignment with the radio axis, contrary to expectations from unified models of active galactic nuclei that rely on orientation effects (e.g., Barthel (1989); Antonucci (1993); Urry & Padovani (1995)). This symmetry of the radiation field, if typical of all quasars, may have important consequences on our understanding of the impact that quasars have on the intergalactic environment (e.g., proximity effect of Ly$`\alpha `$ clouds; Carswell et al. (1982); Murdoch et al. (1986)). ## 4 Summary and Future Directions Our deep H$`\alpha `$ observations of MR 2251–178 reveal a spiral complex of ionized material that extends more or less symmetrically out to $``$ 120 kpc from the quasar. The coarse velocity field derived from our data shows a NE – SW velocity gradient that is opposite to that of the inner line-emitting region. The morphology and kinematics of the nebula cannot be easily explained by an interaction/merger event or a cooling flow. We favor the scenario in which the ionized material is part of a larger neutral envelope that is photoionized by the radiation field of the quasar. Deep H I 21-cm observations are needed to confirm this model. We suspect that significant quantities of ionized gas may be present around luminous quasars, but have remained undetected with standard narrowband imaging techniques. The observations presented herein constitute only 40 minutes of integration, yet reach more than an order of magnitude deeper in flux than previous narrowband imagery. The TTF instrument allows the observer to tune the filter to a very precise bandpass in order to match the expected emission and avoid bright sky features. The TTF therefore promises to greatly impact a number of observational programs, such as those that aim at parameterizing ionized quasar envelopes and quantifying the impact of quasars on the intergalactic environment. SV is grateful for support of this research by NSF/CAREER grant AST-9874973 and by a Cottrell Scholarship awarded by the Research Corporation.